Review Effects of , Prebiotics, and on Human Health

Paulina Markowiak * and Katarzyna Sli´ zewska˙ * Institute of Fermentation Technology and , Department of Biotechnology and Food Sciences, Lodz University of Technology, 90-924 Łód´z,Poland * Correspondence: [email protected] (P.M.); [email protected] (K.S.);´ Tel.: +48-42-6313-481 (P.M. & K.S.)´

Received: 24 July 2017; Accepted: 12 September 2017; Published: 15 September 2017

Abstract: The human is colonised by a complex ecosystem of . Intestinal are not only commensal, but they also undergo a synbiotic co-evolution along with their host. Beneficial intestinal bacteria have numerous and important functions, e.g., they produce various nutrients for their host, prevent infections caused by intestinal pathogens, and modulate a normal immunological response. Therefore, modification of the intestinal microbiota in order to achieve, restore, and maintain favourable balance in the ecosystem, and the activity of microorganisms present in the gastrointestinal tract is necessary for the improved health condition of the host. The introduction of probiotics, prebiotics, or synbiotics into human diet is favourable for the intestinal microbiota. They may be consumed in the form of raw vegetables and fruit, fermented pickles, or dairy products. Another source may be pharmaceutical formulas and functional food. This paper provides a review of available information and summarises the current knowledge on the effects of probiotics, prebiotics, and synbiotics on human health. The mechanism of beneficial action of those substances is discussed, and verified study results proving their efficacy in human are presented.

Keywords: bacteria; prebiotics; synbiotics; human health;

1. Introduction Nowadays, besides the basic role of nutrition consisting in the supply of necessary nutrients for growth and development of the organism, some additional aspects are becoming increasingly important, including the maintenance of health and counteracting diseases. In the world of highly processed food, particular attention is drawn to the composition and safety of consumed products. The quality of food is very important because of, i.e., the problem of food poisoning, obesity, allergy, cardiovascular diseases, and cancer—the plague of the 21st century. Scientific reports point to the health benefits of using probiotics and prebiotics in . The word “probiotic” comes from Greek, and it means “for life”. Most probably, it was Ferdinand Vergin who invented the term “probiotic” in 1954, in his article entitled “Anti-und Probiotika” comparing the harmful effects of and other antibacterial agents on the intestinal microbiota with the beneficial effects (“probiotika”) of some useful bacteria [1]. Some time after that, in 1965, Lilly and Stillwell described probiotics as microorganisms stimulating the growth of other microorganisms [2]. The definition of probiotics has been modified and changed many times. To emphasise their microbial origin, Fuller (1989) stated that probiotics must be viable microorganisms and must exert a beneficial effect on their host [3]. On the other hand, Guarner and Schaafsma (1998) indicated the necessary use of an appropriate dose of probiotic organisms required to achieve the expected effect [4]. The current definition, formulated in 2002 by FAO (Food and Agriculture Organization of the United Nations) and WHO (World Health Organization) working group experts, states that probiotics are “live strains

Nutrients 2017, 9, 1021; doi:10.3390/nu9091021 www.mdpi.com/journal/nutrients Nutrients 2017, 9, 1021 2 of 30 of strictly selected microorganisms which, when administered in adequate amounts, confer a health benefit on the host” [5]. The definition was maintained by the International Scientific Association for Probiotics and Prebiotics (ISAPP) in 2013 [6]. Results of clinical studies confirm the positive effect of probiotics on gastrointestinal diseases (e.g., irritable bowel syndrome, gastrointestinal disorders, elimination of Helicobacter, inflammatory bowel disease, diarrhoeas) and allergic diseases (e.g., atopic dermatitis). Many clinical studies have proven the effectiveness of probiotics for treatment of diseases such as obesity, insulin resistance syndrome, type 2 diabetes, and non-alcoholic fatty liver disease. Furthermore, the positive effects of probiotics on human health have been demonstrated by increasing the body’s immunity (immunomodulation). Scientific reports also show the benefits of the prophylactic use of probiotics in different types of cancer and side effects associated with cancer. Many clinical studies have proven the effectiveness of probiotics, and recommended doses of probiotics are those that have been used in a particular case. Keep in mind that how probiotics work may depend on the strain, dose, and components used to produce a given probiotic product. In 1995, prebiotics were defined by Gibson and Roberfroid as non-digested food components that, through the stimulation of growth and/or activity of a single type or a limited amount of microorganisms residing in the gastrointestinal tract, improve the health condition of a host [7]. In 2004, the definition was updated and prebiotics were defined as selectively fermented components allowing specific changes in the composition and/or activity of microorganisms in the gastrointestinal tract, beneficial for host’s health and wellbeing [8]. Finally, in 2007, FAO/WHO experts described prebiotics as a nonviable food component that confers a health benefit on the host associated with modulation of the microbiota [9]. Prebiotics may be used as an alternative to probiotics or as an additional support for them. However different prebiotics will stimulate the growth of different indigenous gut bacteria. Prebiotics have enormous potential for modifying the gut microbiota, but these modifications occur at the level of individual strains and species and are not easily predicted a priori. There are many reports on the beneficial effects of prebiotics on human health. High potential is attributed to the simultaneous use of probiotics and prebiotics. In 1995, Gibson and Roberfroid introduced the term “synbiotic” to describe a combination of synergistically acting probiotics and prebiotics [7]. A selected component introduced to the gastrointestinal tract should selectively stimulate growth and/or activate the metabolism of a physiological intestinal microbiota, thus conferring beneficial effect to the host’s health [10]. As the word “synbiotic” implies synergy, the term should be reserved for those products in which a component selectively favours a probiotic [11]. The principal purpose of that type of combination is the improvement of survival of probiotic microorganisms in the gastrointestinal tract. Synbiotics have both probiotic and prebiotic properties and were created in order to overcome some possible difficulties in the survival of probiotics in the gastrointestinal tract [12]. Therefore, an appropriate combination of both components in a single product should ensure a superior effect, compared to the activity of the probiotic or prebiotic alone [13,14]. The aim of the review was to discuss the mechanisms of action of probiotics, prebiotics, and synbiotics, as well as the current insight into their effect on human health. The selection of probiotic strains, prebiotics, and their respective dosages is crucial in obtaining a therapeutic effect, so separate sections are dedicated to this topic. Further research into the acquisition of new probiotic strains, the selection of probiotics and prebiotics for synbiotics, dose setting, safety of use, and clinical trials documenting the desired health effects is necessary. Effects should be confirmed in properly scheduled clinical trials conducted by independent research centres.

2. Probiotics The knowledge of the beneficial effects of lactic acid fermentation on human health dates back to ancient times. The Bible mentions sour milk several times. Ancient Romans and Greeks knew various Nutrients 2017, 9, 1021 3 of 30 recipes for fermented milk. A specific type of sour milk, called “leben raib”, prepared from buffalo, cow, or goat milk, was consumed in ancient Egypt. A similar “jahurt” was also commonly consumed by people inhabiting the Balkans. In India, fermented milk drinks were known already 800–300 years B.C., and in Turkey in the 8th century. A milk drink called “ajran” was consumed in Central Russia in the 12th century, and “tarho” was consumed in Hungary in the 14th century [15]. A particular interest in lactic acid fermentation was expressed in the beginning of the 20th century by the Russian scientist and immunologist working for the Pasteur Institute in Paris, awarded with the Nobel Prize in medicine for his work on immunology (in 1907), Ilia Miecznikow. Here is a quote from his book “Studies on Optimism”: “with various foods undergoing lactic acid fermentation and consumed raw (sour milk, kefir, sauerkraut, pickles) humans introduced huge amounts of proliferating to their alimentary tracts” [16].

2.1. Selection Criteria and Requirements for Probiotic Strains According to the suggestions of the WHO, FAO, and EFSA (the European Food Safety Authority), in their selection process, probiotic strains must meet both safety and functionality criteria, as well as those related to their technological usefulness (Table1). Probiotic characteristics are not associated with the genus or species of a microorganism, but with few and specially selected strains of a particular species [6]. The safety of a strain is defined by its origin, the absence of association with pathogenic cultures, and the resistance profile. Functional aspects define their survival in the gastrointestinal tract and their immunomodulatory effect. Probiotic strains have to meet the requirements associated with the technology of their production, which means they have to be able to survive and maintain their properties throughout the storage and distribution processes [17]. Probiotics should also have documented pro-health effects consistent with the characteristics of the strain present in a marketed product. Review papers and scientific studies on one strain may not be used for the promotion of other strains as probiotics. It has to be considered, as well, that the studies documenting probiotic properties of a particular strain at a tested dose do not constitute evidence of similar properties of a different dose of the same strain. Also, the type of carrier/matrix is important, as it may reduce the viability of a particular strain, thus changing the properties of a product [18,19].

Table 1. Selection criteria of probiotic strains [5,20].

Criterion Required Properties • Human or animal origin. • Isolated from the gastrointestinal tract of healthy individuals. • History of safe use. • Precise diagnostic identification (phenotype and genotype traits). Safety • Absence of data regarding an association with infective disease. • Absence of the ability to cleave bile acid salts. • No adverse effects. • Absence of genes responsible for antibiotic resistance localised in non-stable elements. • Competitiveness with respect to the microbiota inhabiting the intestinal ecosystem. • Ability to survive and maintain the metabolic activity, and to grow in the target site. • Resistance to bile salts and enzymes. • Resistance to low pH in the stomach. • Competitiveness with respect to microbial species inhabiting the intestinal ecosystem Functionality (including closely related species). • Antagonistic activity towards pathogens (e.g., H. pylori, Salmonella sp., Listeria monocytogenes, difficile). • Resistance to bacteriocins and acids produced by the endogenic intestinal microbiota. • Adherence and ability to colonise some particular sites within the host organism, and an appropriate survival rate in the gastrointestinal system. • Easy production of high biomass amounts and high productivity of cultures. • Viability and stability of the desired properties of probiotic bacteria during the fixing process (freezing, freeze-drying), preparation, and distribution of probiotic products. Technological • High storage survival rate in finished products (in aerobic and micro-aerophilic conditions). usability • Guarantee of desired sensory properties of finished products (in the case of the food industry). • Genetic stability. • Resistance to bacteriophages. Nutrients 2017, 9, 1021 4 of 30

2.2. Probiotic Microorganisms Probiotic products may contain one or more selected microbial strains. Human probiotic microorganisms belong mostly to the following geni: , , and Lactococus, Streptococcus, Enterococcus. Moreover, strains of Gram-positive bacteria belonging to the genus Bacillus and some yeast strains belonging to the genus Saccharomyces are commonly used in probiotic products [21]. Probiotics are subject to regulations contained in the general food law, according to which they should be safe for human and animal health. In the USA, microorganisms used for consumption purposes should have the GRAS (Generally Regarded As Safe) status, regulated by the FDA (Food and Drug Administration). In Europe, EFSA introduced the term of QPS (Qualified Presumption of Safety). The QPS concept involves some additional criteria of the safety assessment of bacterial supplements, including the history of safe usage and absence of the risk of acquired resistance to antibiotics [22,23]. Table2 presents probiotic microorganisms contained in pharmaceutical products and as food additives.

Table 2. Probiotic microorganisms used in human nutrition [24–26].

Type Lactobacillus Type Bifidobacterium Other Lactic Acid Bacteria Other Microorganisms L. acidophilus (a),* L. amylovorus (b),* L. casei (a),(b),* B. adolescentis (a) L. gasseri (a),* B. animalis (a),* Bacillus clausii (a),* Enterococcus faecium (a) L. helveticus (a),* B. bifidum (a) Escherichia coli Nissle 1917 (a) Lactococcus lactis (b),* L. johnsonii (b),* B. breve (b) Saccharomyces cerevisiae Streptococcus thermophilus (a),* L. pentosus (b),* B. infantis (a) (boulardi) (a),* L. plantarum (b),* B. longum (a),* L. reuteri (a),* L. rhamnosus (a),(b),* (a) Mostly as pharmaceutical products; (b) mostly as food additives; * QPS (Qualified Presumption of Safety) microorganisms.

2.3. of Probiotics A significant progress has been observed lately in the field of studies on probiotics, mostly in terms of the selection and characteristics of individual probiotic cultures, their possible use, and their effect on health. Probiotics have numerous advantageous functions in human organisms. Their main advantage is the effect on the development of the microbiota inhabiting the organism in the way ensuring proper balance between pathogens and the bacteria that are necessary for a normal function of the organism [27,28]. Live microorganisms meeting the applicable criteria are used in the production of functional food and in the preservation of food products. Their positive effect is used for the restoration of natural microbiota after antibiotic therapy [29,30]. Another function is counteracting the activity of pathogenic intestinal microbiota, introduced from contaminated food and environment. Therefore, probiotics may effectively inhibit the development of pathogenic bacteria, such as Clostridium perfringens [31], Campylobacter jejuni [32], Salmonella Enteritidis [33], Escherichia coli [34], various species of Shigella [35], Staphylococcus [36], and Yersinia [37], thus preventing food poisoning. A positive effect of probiotics on digestion processes, treatment of food allergies [38,39], candidoses [40], and dental caries [41] has been confirmed. Probiotic microorganisms such as [42], Lactobacillus reuteri [43], Bifidobacterium adolescentis, and Bifidobacterium pseudocatenulatum [44] are natural producers of B group (B1, B2, B3, B6, B8, B9, B12). They also increase the efficiency of the immunological system, enhance the absorption of vitamins and compounds, and stimulate the generation of organic acids and amino acids [18,45–47]. Probiotic microorganisms may also be able to produce enzymes, such as esterase, lipase, and co-enzymes A, Q, NAD, and NADP. Some products of probiotics’ metabolism may also show antibiotic (acidophiline, bacitracin, lactacin), anti-cancerogenic, and immunosuppressive properties [45,48–50]. Molecular and genetic studies allowed the determination of the basics of the beneficial effect of probiotics, involving four mechanisms: Nutrients 2017, 9, 1021 5 of 30

(1) Antagonism through the production of antimicrobial substances [51]; (2) Competition with pathogens for adhesion to the epithelium and for nutrients [52]; (3) Immunomodulation of the host [53]; (4) Inhibition of bacterial toxin production [54].

The first two mechanisms are directly associated with their effect on other microorganisms. Those mechanisms are important in prophylaxis and treatment of infections, and in the maintenance of balance of the host’s intestinal microbiota. The ability of probiotic strains to co-aggregate, as one of their mechanisms of action, may lead to the formation of a protective barrier preventing pathogenic bacteria from the colonisation of the epithelium [27]. Probiotic bacteria may be able to adhere to epithelial cells, thus blocking pathogens. That mechanism exerts an important effect on the host’s health condition. Moreover, the adhesion of probiotic microorganisms to epithelial cells may trigger a signalling cascade, leading to immunological modulation. Alternatively, the release of some soluble components may cause a direct or indirect (through epithelial cells) activation of immunological cells. This effect plays an important role in the prevention and treatment of contagious diseases, as well as in chronic inflammation of the alimentary tract or of a part thereof [28]. There are also suggestions of a possible role of probiotics in the elimination of cancer cells [55]. Results of in vitro studies indicate the role of low-molecular-weight substances produced by probiotic microorganisms (e.g., hydroperoxide and short-chain fatty acids) in inhibiting the replication of pathogens [28]. For example, Lactobacillus genus bacteria may be able to produce bacteriocins, including low-molecular-weight substances (LMWB—antibacterial peptides), as well as high-molecular-weight ones (class III bacteriocins), and some antibiotics. Probiotic bacteria (e.g., Lactobacillus and Bifidobacterium) may produce the so-called de-conjugated bile acids (derivatives of bile acids), demonstrating stronger antibacterial effect than the bile salts produced by their host [28,56]. Further studies are necessary to explain the mechanism of acquiring resistance to their own metabolites by Lactobacillus genus bacteria. The essential for nearly all bacteria, except for lactic acid bacteria, is . It turns out that Lactobacillus bacteria do not need iron in their natural environment, which may be their crucial advantage over other microorganisms [57]. Lactobacillus delbrueckii affects the function of other microbes by binding iron hydroxide to its cellular surface, thus making it unavailable to other microbes [58]. The immunomodulatory effect of the intestinal microbiota, including probiotic bacteria, is based on three, seemingly contradictory phenomena [53,59]: (1) Induction and maintenance of the state of immunological tolerance to environmental antigens (nutritional and inhalatory); (2) Induction and control of immunological reactions against pathogens of bacterial and viral origin; (3) Inhibition of auto-aggressive and allergic reactions. Probiotic-induced immunological stimulation is also manifested by the increased production of immunoglobulins, enhanced activity of macrophages and lymphocytes, and stimulation of γ-interferon production. Probiotics may influence the congenital and acquired immunological system through metabolites, components of the cellular wall, and DNA, recognised by specialised cells of the host (e.g., those equipped with receptors) [28]. The principal host cells that are important in the context of the immune response are intestinal epithelial cells and intestinal immune cells. Components of the cellular wall of lactic acid bacteria stimulate the activity of macrophages. Those, in turn, are able to destroy microbes rapidly by the increased production of free oxygen radicals and lysosomal enzymes. Probiotic bacteria are also able to stimulate the production of cytokines by immunocompetent cells of the gastrointestinal tract [60]. On the other hand, the immunological activity of yeast is associated with the presence of glucans in their cellular wall. Those compounds stimulate the response of the reticuloendothelial system [61]. The last of the abovementioned probiotic effects—inhibition of the production of bacterial toxins—is based on actions leading to toxin inactivation and help with the removal of toxins from Nutrients 2017, 9, 1021 6 of 30 the body. Help in detoxification from the body can take place by adsorption (some strains can bind toxins to their cell wall and reduce the intestinal absorption of toxins), but can also result from the metabolism of mycotoxins (e.g., aflatoxin) by microorganisms [62–64]. However, not all probiotics exhibit detoxifying properties, as it is a strain-related characteristic. Studies should therefore be conducted to select strains with such characteristics. The effectiveness of some probiotics in combating diarrhoea is probably associated with their ability to protect the host from toxins. The reduction of metabolic reactions leading to the production of toxins is also associated with the stimulation of pathways leading to the production of native enzymes, vitamins, and antimicrobial substances [28]. Gut microbiota play a significant role in host metabolic processes (e.g., the regulation of cholesterol absorption, blood pressure (BP), and glucose metabolism), and recent metagenomic surveys have revealed that they are involved in host immune modulation and that they influence host development and physiology (organ development) [65–67]. Nutritional programming to manipulate the composition of the intestinal microbiota through the administration of probiotics continues to receive much attention for the prevention or attenuation of the symptoms of metabolic-related diseases. Currently, studies are exploring the potential for expanded uses of probiotics for improving health conditions in metabolic disorders that increase the risk of developing cardiovascular diseases such as hypertension. Further investigations are required to evaluate the targeted and effective use of the wide variety of probiotic strains in various metabolic disorders to improve the overall health status of the host [65]. In order to confirm the beneficial role of probiotics in improving cardiovascular health and in the reduction of BP, more extensive studies are needed to understand the mechanisms underlying probiotic action. Most probably, all of the abovementioned mechanisms of probiotic action have an effect on the protection against infections, cancer, and the stabilization of balance of the host’s intestinal microbiota. However, it seems unlikely that each of the probiotic microorganisms has properties of all four aspects simultaneously and constitutes a universal remedy to multiple diseases. An important role in the action of probiotics is played by species- and strain-specific traits, such as: cellular structure, cell surface, size, metabolic properties, and substances secreted by microorganisms. The use of a combination of probiotics demonstrating various mechanisms of action may provide enhanced protection offered by a bio-therapeuticNutrients 2017 product, 9, 1021 [68]. Figure1 summarises the mechanisms and effects of action of7 probiotics.of 28

Figure 1. Mechanisms of action of synbiotics and their effects. Figure 1. Mechanisms of action of synbiotics and their effects.

2.4. Probiotics for Humans In the face of widespread diseases and ageing societies, the use of knowledge on microbiocenosis of the gastrointestinal tract and on the beneficial effect of probiotic bacteria is becoming increasingly important. The consumption of pre-processed food (fast food), often containing excessive amounts of fat and insufficient amounts of vegetables, is another factor of harmful modification of human intestinal microbiota. There is currently no doubt about the fact that the system of intestinal microorganisms and its desirable modification with probiotic formulas and products may protect people against enteral problems, and influence the overall improvement of health. Probiotics may be helpful in the treatment of inflammatory enteral conditions, including , Crohn’s disease, and non-specific ileitis. The aetiology of those diseases is not completely understood, but it is evident that they are associated with chronic and recurrent infections or inflammations of the intestine. Clinical studies have demonstrated that probiotics lead to the remission of ulcerative colitis, but no positive effect on Crohn’s disease has been observed [69,70]. Numerous studies assessed the use of probiotics in the treatment of lactose intolerance [71,72], irritable bowel syndrome, and the prevention of [73] and peptic ulcers [74]. Considering their role in the inhibition of some bacterial enzymes, probiotics may reduce the risk of colorectal carcinoma in animals. However, the same effect in humans has not been confirmed in clinical trials [75]. On the other hand, a positive effect on the urogenital system (prevention and treatment of Urinary Tract Infections (UTIs) and bacterial vaginitis) constitutes an excellent example of the benefits associated with the use of probiotics [76–78]. There were attempts to apply probiotics to pregnant women and neonates in order to prevent allergic diseases such as atopic dermatitis. However, the scope of action is controversial in this kind of case [79]. There is evidence that the consumption of probiotics-containing dairy products results in the reduction of blood cholesterol, which may be helpful in the prevention of obesity, diabetes, cardiovascular diseases, and cerebral stroke [80]. The reduction of cholesterol level achieved due to probiotics is less pronounced compared to the effect of pharmaceutical agents, but leads to a significant minimisation of side effects [80]. Other studies confirmed the effect of the probiotic formula VSL#3 and of the Oxalobacter formigenes bacterial

Nutrients 2017, 9, 1021 7 of 30

2.4. Probiotics for Humans In the face of widespread diseases and ageing societies, the use of knowledge on microbiocenosis of the gastrointestinal tract and on the beneficial effect of probiotic bacteria is becoming increasingly important. The consumption of pre-processed food (fast food), often containing excessive amounts of fat and insufficient amounts of vegetables, is another factor of harmful modification of human intestinal microbiota. There is currently no doubt about the fact that the system of intestinal microorganisms and its desirable modification with probiotic formulas and products may protect people against enteral problems, and influence the overall improvement of health. Probiotics may be helpful in the treatment of inflammatory enteral conditions, including ulcerative colitis, Crohn’s disease, and non-specific ileitis. The aetiology of those diseases is not completely understood, but it is evident that they are associated with chronic and recurrent infections or inflammations of the intestine. Clinical studies have demonstrated that probiotics lead to the remission of ulcerative colitis, but no positive effect on Crohn’s disease has been observed [69,70]. Numerous studies assessed the use of probiotics in the treatment of lactose intolerance [71,72], irritable bowel syndrome, and the prevention of colorectal cancer [73] and peptic ulcers [74]. Considering their role in the inhibition of some bacterial enzymes, probiotics may reduce the risk of colorectal carcinoma in animals. However, the same effect in humans has not been confirmed in clinical trials [75]. On the other hand, a positive effect on the urogenital system (prevention and treatment of Urinary Tract Infections (UTIs) and bacterial vaginitis) constitutes an excellent example of the benefits associated with the use of probiotics [76–78]. There were attempts to apply probiotics to pregnant women and neonates in order to prevent allergic diseases such as atopic dermatitis. However, the scope of action is controversial in this kind of case [79]. There is evidence that the consumption of probiotics-containing dairy products results in the reduction of blood cholesterol, which may be helpful in the prevention of obesity, diabetes, cardiovascular diseases, and cerebral stroke [80]. The reduction of cholesterol level achieved due to probiotics is less pronounced compared to the effect of pharmaceutical agents, but leads to a significant minimisation of side effects [80]. Other studies confirmed the effect of the probiotic formula VSL#3 and of the Oxalobacter formigenes bacterial strain on the elimination of oxalates with urine, which may potentially reduce the risk of urolithiasis [81]. Studies on animals demonstrated that orally administered Lactobacillus acidophilus induces expression of µ-opioid and cannabinoid receptors in intestinal cells and mediate analgesic functions in the intestine, and that the observed effect is comparable to the effect of morphine [82]. However, the effect has not been demonstrated in humans. There are many reports on the application of probiotics in the treatment of diarrhoea. The application of Saccharomyces boulardii yeast to patients with acute, watery diarrhoea resulted in the cure and reduced frequency of that type of complaints in two subsequent months [83]. The efficacy of probiotic strains in the therapy of nosocomial, non-nosocomial, and viral diarrhoeas has also been documented. It turns out that probiotics may increase the amount of IgA antibodies, which leads to the arrest of a viral infection [84]. Antibiotic-associated diarrhoea (AAD) is a common complication of most antibiotics and Clostridium difficile disease (CDD), which also is incited by antibiotics, and is a leading cause of nosocomial outbreaks of diarrhoea and colitis. The use of probiotics for these two related diseases remains controversial. A variety of different types of probiotics show promise as effective therapies for these two diseases. Using meta-analyses, three types of probiotics (Saccharomyces boulardii, GG, and probiotic mixtures) significantly reduced the development of antibiotic-associated diarrhoea. Only S. boulardii was effective for CDD [85]. Studies performed in a foster home in Helsinki (Finland) demonstrated that the regular use of Lactobacillus rhamnosus GG in the form of a probiotic resulted in a reduced number of respiratory tract infections [86]. Other studies demonstrated that the application of a diet depleted of fermented foods caused a reduction of congenital immunological response, as well as a significant reduction of stool Lactobacillus count and of the stool amount of short-chain fatty acids. Moreover, the reduction Nutrients 2017, 9, 1021 8 of 30 of phagocytic activity of leukocytes was observed after two weeks of the diet, which could have a negative impact on the organism’s ability to protect against infections [87]. The effect of a fermented product containing Lactobacillus gasseri CECT5714 and Lactobacillus coryniformis CECT5711 strains on blood and stool parameters was studied in a randomised, double-blind trial on 30 healthy volunteers. No negative effects were observed in the group of subjects receiving the probiotic strains. Some positive effects were observed, including: the production of short-chain fatty acids, humidity, frequency and volume of stools, and subjective improvement of intestinal function [88]. Studies by Alvaro et al. (2007) demonstrated a significant reduction of Enterobacteriaceae count and increased galactosidase activity in the alimentary tract of yoghurt consumers, compared to those who did not eat yoghurt [89]. Table3 lists the results of studies focusing on the effect of probiotics on human health. There are examples of clinical trials during which the probiotics group received the probiotic prophylactically or in addition to the standard therapy.

Table 3. Examples of clinical trials regarding the effect of probiotics on human health.

Time of References Subjects Microorganism Main Outcome Administration Obesity 50 obese Increase in the ratios of Bacteroides, Prevotellae, [90] L. salivarius Ls-33 12 weeks adolescents and Porphyromonas. 50 adolescents [91] L. salivarius Ls-33 12 weeks No effect. with obesity 87 subjects with Reduction in BMI, waist, abdominal VFA, and [92] L. gasseri SBT2055 12 weeks high BMI hip circumference. 210 adults with [93] L. gasseri SBT2055 12 weeks Reduction in BMI and arterial BP values. large VFA 40 adults with [94] L. plantarum 3 weeks Reduction in BMI and arterial BP values. obesity L. acidophilus La5, 75 subjects with Changes in gene expression in PBMCs as well as BMI, [95–97] B. lactis Bb12, L. 8 weeks high BMI fat percentage, and leptin levels. casei DN001 70 overweight and E. faecium and 2, S. Reduction in body weight, systolic BP, LDL-C, and [98] 8 weeks obese subjects thermophilus strains increase in fibrinogen levels. Bifidobacterium, 60 overweight Improvement in lipid profile, insulin sensitivity, and [99] Lactobacillus, S. 6 weeks subjects decrease in CRP. thermophilus 58 obese PM [100] L. paracasei N19 6 weeks No effect. women L. acidophilus La5, 156 overweight Reduction in fasting glucose concentration and increase [101] B. animalis subsp. 6 weeks adults in HOMA-IR. lactis Bb12 Insulin resistance syndrome 28 patients [102] L. casei Shirota 12 weeks No effect. with IRS 30 patients [103] L. casei Shirota 12 weeks Significant reduction in the VCAM-1 level. with IRS 24 PM women Glucose and homocysteine levels were [104] L. plantarum 12 weeks with IRS significantly reduced. Type 2 diabetes 40 patients [105] L. planatarum A7 8 weeks Decreased methylation process, SOD, and 8-OHDG. with T2D L. acidophilus La-5, 45 patients Significant difference between groups concerning mean [106] B. animalis subsp. 6 weeks with T2D changes of HbA1c, TC, and LDL-C. lactis BB-12 Nutrients 2017, 9, 1021 9 of 30

Table 3. Cont.

Time of References Subjects Microorganism Main Outcome Administration L. acidophilus La-5, 44 patients Increased HDL-C levels and decreased [107] B. animalis subsp. 8 weeks with T2D LDL-C/HDL-C ratio. lactis BB-12 64 patients L. acidophilus La5, [108] 6 weeks Reduced fasting blood glucose and antioxidant status. with T2D B. lactis Bb12 60 patients L. acidophilus La5, [109] 6 weeks TC and LDL-C improvement. with T2D B. lactis Bb12 L. acidophilus [110] 45 males with T2D 4 weeks No effect. NCFM Non-alcoholic fatty liver disease 20 obese children [111] L. rhamnosus GG 8 weeks Decreased ALT and PG-PS IgAg antibodies. with NAFLD 28 adult L. bulgaris, S. [112] individuals with 12 weeks Decreased ALT and γ-GTP levels. thermophilus NAFLD L.acidophilus La5, B. 72 patients with [113] breve subsp. lactis 8 weeks Reduced serum levels of ALT, ASP, TC, and LDL-C. NAFLD Bb12 Bifidobacterium, 44 obese children Improved fatty liver severity, decreased BMI, and [114] Lactobacillus, S. 16 weeks with NAFLD increased GLP1/aGLP1. thermophilus Irritable bowel syndrome (IBS), gastrointestinal disorders, elimination of Helicobacter, inflammatory bowel disease (IBD), diarrhoeas 59 adults infected L. acidophilus La5, [115] 6 weeks Inhibitory effect against Helicobacter pylori. with H. pylori B. lactis Bb12 16 patients infected Inhibited growth of Helicobacter pylori (by 64% in the [116] L. casei Shirota 6 weeks with H. pylori probiotic group, and by 33% in the control). 269 children with Diarrhoea was less common in children receiving otitis media and/or S. cerevisiae [117] No data probiotic yeast (7.5%) compared to those receiving respiratory tract (boulardii) placebo (23%). No negative side effects were observed. infections 77 patients with Remission in 42.9% of patients in the probiotic group, [118] Probiotic VSL#3 12 weeks ulcerative colitis and in 15.7% of patients in the placebo group. 90 breastfed Elimination of pain and symptoms associated with L. reuteri ATCC [119] neonates with 6 months intestinal colic already after one week of the use of 55730 intestinal colic the probiotic. Atopic dermatitis women—from 512 pregnant 35 weeks gestation women and 474 L. rhamnosus until 6 months if Substantially reduced the cumulative prevalence of [120] their newborn HN001 breastfeeding, eczema in infants. infants infants—from birth to 2 years 53 children with L. fermentum VRI [121] moderate of severe 8 weeks Reduction in SCORAD. 033 PCCTM atopic dermatitis 156 mothers of Mothers—the last 6 high-risk children weeks of Significantly reduction eczema in high-risk for a (i.e., positive family B. bifidum, B. lactis, [122] pregnancy, minimum of 2 years provided that the probiotic was history of allergic L. lactis offspring—12 administered to the infant within 3 months of birth. disease) and their months offspring 50 children B. animalis subsp Significant reduction in the severity of AD with an [123] 8 weeks with AD lactis improved ration of IFN-γ and IL-10. Nutrients 2017, 9, 1021 10 of 30

Table 3. Cont.

Time of References Subjects Microorganism Main Outcome Administration Alleviation of lactose intolerance S. lactis, 15 healthy, L. plantarum, free-living adults S. cremoris, L. casei, [124] 1 day Improved lactose digestion and tolerance. with lactose S. diacetylactis, maldigestion S. florentinus, L. cremoris B. animalis subsp. animalis IM386 (DSM 26137), [125] 44 patients 6 weeks A significant lowering effect on diarrhoea and flatulence. L. plantarum MP2026 (DSM 26329) Different types of cancer and side effects associated with cancer L. plantarum 100 patients with CGMMCC No Improvement in the integrity of gut mucosal barrier and [126] colorectal 1258, L. acidophilus 16 days decrease in infections complications. carcinoma LA-11, B. longum BL-88 63 patients with diarrhoea during L. acidophilus, Reduction in incidence of diarrhoea and better [127] 7 weeks radiotherapy in B. bifidum stool consistency. cervical cancer 150 patients Patients had less grade 4 or 4 diarrhoea, less abdominal [128] diagnosed with L. rhamnosus 573 24 weeks discomfort, needed less hospital care, and had fewer colorectal cancer chemo dose reductions due to bowel toxicity. Abbreviations: AD—atopic dermatitis; ALT—alanine amino transferase; ASP—aspartate amino transferase; BMI—body mass index; BP—blood pressure; CRP—C-reactive protein; γ-GTP—γ-glutamyltranspeptidase; GLP1—glucagon-like peptide 1; HDL-C—high-density lipoprotein cholesterol; HOMA-IR—homeostasis model assessment of insulin resistance; IL-10—interleukin 10; LDL-C—low-density lipoprotein cholesterol; NAFLD—non-alcoholic fatty liver disease; PBMC—peripheral blood mononuclear cell; PM—postmenopausal; SCORAD—SCORing Atopic Dermatitis; SOD—superoxide dismutase, sVCAM-1—soluble vascular cell adhesion molecule-1; TC—total cholesterol; T2D—type 2 diabetes; VFA—visceral fat area; 8-OHDG—8-hydroxy-20-deoxyguanosine.

3. Prebiotics Different prebiotics will stimulate the growth of different indigenous gut bacteria. Prebiotics have enormous potential for modifying the gut microbiota, but these modifications occur at the level of individual strains and species and are not easily predicted a priori. Furthermore, the gut environment, especially pH, plays a key role in determining the outcome of interspecies competition. Both for reasons of efficacy and of safety, the development of prebiotics intended to benefit human health has to take account of the highly individual species profiles that may result [129]. Fruit, vegetables, cereals, and other edible plants are sources of constituting potential prebiotics. The following may be mentioned as such potential souces: tomatoes, artichokes, , , berries, , , chicory, green vegetables, legumes, as well as oats, linseed, , and wheat [130]. Some artificially produced prebiotics are, among others: lactulose, , , maltooligosaccharides, cyclodextrins, and lactosaccharose. Lactulose constitutes a significant part of produced (as much as 40%). , such as and oligofructose, are believed to be the most used and effective in relation to many species of probiotics [131].

3.1. Prebiotic Selection Criteria According to Wang (2009), there are five basic criteria for the classification of food components such as prebiotics (Figure2)[ 132]. The first criterion assumes that prebiotics are not digested (or just partially digested) in the upper segments of the alimentary tract. As a consequence, they reach the colon, where they are selectively fermented by potentially beneficial bacteria (a requirement of the Nutrients 2017, 9, 1021 11 of 28

Nutrientscolon, where2017, 9, 1021they are selectively fermented by potentially beneficial bacteria (a requirement 11of of the 30 second criterion) [133]. The fermentation may lead to the increased production or a change in the secondrelative criterion) abundance [133 of]. different The fermentation short-chain may fatty lead acids to the (SCFAs), increased increased production stool ormass, a change a moderate in the relativereduction abundance of colonic ofpH, different reduction short-chain of nitrous fattyend products acids (SCFAs), and faecal increased enzymes, stool and mass, an improvement a moderate reductionof the immunological of colonic pH, system reduction [134], of nitrouswhich is end benefi productscial for and the faecal host enzymes, (the requirement and an improvement of the third ofcriterion). the immunological Selective stimulation system [134 of], whichgrowth is and/or beneficial activity for the of hostthe intestinal (the requirement bacteria of potentially the third criterion).associated with Selective health stimulation protection and of growth wellbeing and/or is considered activity another of the intestinal criterion [8]. bacteria The last potentially criterion associatedof the classification with health assumes protection that anda prebiotic wellbeing must is considered be able to anotherwithstand criterion food processing [8]. The last conditions criterion ofand the remained classification unchanged, assumes thatnon-degraded, a prebiotic must or bechemically able to withstand unaltered food and processing available conditions for bacterial and remainedmetabolism unchanged, in the intestine non-degraded, [132]. Huebner or chemically et al unaltered. (2008) tested and available several forcommercially bacterial metabolism available inprebiotics the intestine using [ 132various]. Huebner processing et al. conditions. (2008) tested They several found commercially no significant available changes prebiotics of the prebiotic using variousactivity processingof the tested conditions. substances They in foundvarious no proce significantssing conditions changes of [135]. the prebiotic Meanwhile, activity Ze ofet theal. tested(2012) substancesshowed that in it various was possible processing to alter conditions the ability [135 of]. gut Meanwhile, bacteria by Ze utilising et al. (2012) showed in vitro that [136]. it wasThe possiblestructure to alterof prebiotics the ability ofshould gut bacteria be appropri by utilisingately starch documented,in vitro [136 and]. The components structure of prebioticsused as shouldpharmaceutical be appropriately formulas, documented, food, or feed and additives components should used be as relatively pharmaceutical easy to formulas,obtain at food,an industrial or feed additivesscale [137]. should be relatively easy to obtain at an industrial scale [137].

Prebiotic selection criteria

Resistance to Stability in various digestion in the upper Selective stimulation Fermentation by Beneficial effect on food/feed processing sections of the of growth of intestinal microbiota. host's health. conditions. alimentary tract. probiotics.

FigureFigure 2.2. RequirementsRequirements forfor potentialpotential prebioticsprebiotics [[132,138].132,138].

PrebioticsPrebiotics may be be used used as as an an alternative alternative to probiotics to probiotics or as or an as additional an additional support support for them. for Long- them. Long-termterm stability stability during during the shelf-life the shelf-life of food, of food, drinks drinks,, and and feed, feed, resistance resistance to to processing, processing, and physicalphysical andand chemicalchemical propertiesproperties thatthat exhibitexhibit aa positivepositive effecteffect onon thethe flavourflavour andand consistenceconsistence ofof productsproducts maymay promotepromote prebioticsprebiotics asas aa competitioncompetition toto probiotics.probiotics. Additionally,Additionally, resistanceresistance toto acids,acids, proteases,proteases, andand bilebile saltssalts presentpresent inin thethe gastrointestinalgastrointestinal tracttract maymay bebe consideredconsidered asas otherother favourablefavourable propertiesproperties ofof prebiotics.prebiotics. PrebioticPrebiotic substancessubstances selectivelyselectively stimulatestimulate microorganismsmicroorganisms presentpresent inin thethe host’shost’s intestinalintestinal ecosystem,ecosystem, thusthus eliminatingeliminating thethe needneed forfor competitioncompetition withwith bacteria.bacteria. StimulationStimulation ofof thethe intestinalintestinal microbiotamicrobiota byby prebioticsprebiotics determinesdetermines theirtheir fermentationfermentation activity, simultaneouslysimultaneously influencinginfluencing thethe SCFA level,level, whichwhich confers a a health health benefit benefit on on the the host host [139,140]. [139,140]. Moreover, Moreover, prebiotics prebiotics cause cause a reduction a reduction of ofintestinal intestinal pH pH and and maintain maintain the the osmotic osmotic retention retention of ofwater water in inthe the bowel bowel [134]. [134 However,]. However, it should it should be beconsidered considered that that an an overdose overdose of of prebiotics prebiotics may may le leadad to to flatulence and diarrhoea—these effects areare absentabsent inin thethe case case of of excessive excessive consumption consumption of probiotics.of probiotics. Prebiotics Prebiotics may may be consumedbe consumed on a on long-term a long- basisterm andbasis for and prophylactic for prophylactic purposes. purposes. Moreover, Moreover, when used when at correct used doses, at correct they dodoses, notstimulate they do anynot adversestimulate effects, any adverse such as effects, diarrhoea, such susceptibility as diarrhoea, tosusceptibility UV light, or to hepatic UV light, injuries or hepatic caused injuries by antibiotics. caused Prebioticby antibiotics. substances Prebiotic are not substances allergenic are and not do alle notrgenic proliferate and thedo abundancenot proliferate of antibiotic-resistance the abundance of genes.antibiotic-resistance Of course, the genes. effect ofOf the course, elimination the effect of selectedof the elimination pathogens of achieved selected bypathogens the useof achieved prebiotics by maythe use be inferiorof prebiotics to antibiotics, may be inferior but the propertiesto antibiotic mentioneds, but the aboveproperties make mentioned them a natural above substitute make them for antibioticsa natural substitute [134]. for antibiotics [134].

3.2.3.2. PrebioticPrebiotic SubstancesSubstances TheThe majoritymajority ofof identifiedidentified prebioticsprebiotics areare carbohydratescarbohydrates ofof various molecular structures, naturallynaturally occurringoccurring inin humanhuman andand animalanimal diets.diets. TheThe physiologicalphysiological propertiesproperties ofof potentialpotential prebioticsprebiotics determinedetermine theirtheir beneficialbeneficial effecteffect onon thethe host’shost’s health.health. PrebioticsPrebiotics maymay bebe classifiedclassified accordingaccording toto thosethose propertiesproperties asas [[134]:134]:  not digested (or only partially digested);

Nutrients 2017, 9, 1021 12 of 30

• not digested (or only partially digested); • not absorbed in the small intestine; • poorly fermented by bacteria in the oral cavity; • well fermented by seemingly beneficial intestinal bacteria; • poorly fermented by potential pathogens in the bowel.

Carbohydrates, such as dietary fibre, are potential prebiotics. Prebiotic and dietary fibre are terms used alternatively for food components that are not digested in the gastrointestinal tract. A significant difference between those two terms is that prebiotics are fermented by strictly defined groups of microorganisms, and dietary fibre is used by the majority of colonic microorganisms [141]. Therefore, considering one of the basic classification criteria, it turns out that using those terms alternatively is not always correct. Prebiotics may be a dietary fibre, but dietary fibre is not always a prebiotic [138]. The following non-starch polysaccharides are considered to be dietary fibre: cellulose, hemicellulose, , gums, substances obtained from marine algae, as well as lactulose, soy oligosaccharides, , fructooligosaccharides, galactooligosaccharides, , and isomaltooligosaccharides. Based on the number of monomers bound together, prebiotics may be classified as: disaccharides, oligosaccharides (3–10 monomers), and polysaccharides. The most promising and fulfilling criteria for the classification of prebiotic substances, as evidenced by in vitro and in vivo studies, are oligosaccharides, including [142,143]: fructooligosaccharides (FOS), galactooligosaccharides (GOS), isomaltooligosaccharides (IMO), xylooligosaccharides (XOS), transgalactooligosaccharides (TOS), and soybean oligosaccharides (SBOS). Also, polysaccharides such as inulin, reflux starch, cellulose, hemicellulose, or may potentially be prebiotics. Examples of prebiotics that are most commonly used in human nutrition are presented in Table4. The use of glucooligosaccharides, glicooligosaccharides, lactitol, izomaltooligosaccharides, stachyose, raffinose, and saccharose as prebiotics requires further studies [144].

Table 4. Examples of prebiotics and synbiotics used in human nutrition [134,145,146].

Human Nutrition Prebiotics Synbiotics FOS GOS Inulin Lactobacillus genus bacteria + inulin XOS Lactobacillus, Streptococcus and Bifidobacterium genus bacteria + FOS Lactitol Lactobacillus, Bifidobacterium, Enterococcus genus bacteria + FOS Lactosucrose Lactobacillus and Bifidobacterium genus bacteria + oligofructose Lactulose Lactobacillus and Bifidobacterium genus bacteria + inulin Soy oligosaccharides TOS Abbreviations: FOS—fructooligosaccharides; GOS—galactooligosaccharides; TOS—transgalactooligosaccharides; XOS—xylooligosaccharides.

3.3. Mechanism of Action of Prebiotics Prebiotics are present in natural products, but they may also be added to food. The purpose of these additions is to improve their nutritional and health value. Some examples are: inulin, fructooligosaccharides, lactulose, and derivatives of galactose and β-glucans. Those substances may serve as a medium for probiotics. They stimulate their growth, and contain no microorganisms. Figure2 presents the principal mechanisms of prebiotic action and some of their effects on the host’s health. Prebiotics are not digested by host enzymes and reach the colon in a practically unaltered form, where they are fermented by saccharolytic bacteria (e.g., Bifidobacterium genus). The consumption of prebiotics largely affects the composition of the intestinal microbiota and its metabolic activity [147]. This is due to the modulation of lipid metabolism, enhanced absorbability of , effect on Nutrients 2017, 9, 1021 13 of 30 the immunological system, and modification of the bowel function [147]. It is highly probable that providing an energy source that only specific species in the microbiota can utilize has a greater impact on microbiota composition and metabolism than these other factors. The molecular structure of prebiotics determines their physiological effects and the types of microorganisms that are able to use them as a source of carbon and energy in the bowel [134]. It was demonstrated that, despite the variety of carbohydrates that exhibit the prebiotic activity, the effect of their administration is an increased count of beneficial bacteria, mostly of the Bifidobacterium genus [148,149]. The mechanism of a beneficial effect of prebiotics on immunological functions remains unclear. Several possible models have been proposed [150]:

(1) Prebiotics are able to regulate the action of hepatic lipogenic enzymes by influencing the increased production of short-chain fatty acids (SCFAs), such as propionic acid. (2) The production of SCFAs (especially of ) as a result of fermentation was identified as a modulator of histone acetylation, thus increasing the availability of numerous genes for transcription factors. (3) The modulation of mucin production. (4) It was demonstrated that FOS and several other prebiotics cause an increased count of lymphocytes and/or leukocytes in gut-associated lymphoid tissues (GALTs) and in peripheral blood. (5) The increased secretion of IgA by GALTs may stimulate the phagocytic function of intra-inflammatory macrophages.

The main aim of prebiotics is to stimulate the growth and activity of beneficial bacteria in the gastrointestinal tract, which confers a health benefit on the host. Through mechanisms including antagonism (the production of antimicrobial substances) and competition for epithelial adhesion and for nutrients, the intestinal microbiota acts as a barrier for pathogens. Final products of metabolism are mostly SCFAs, namely: acetic acid, butyric acid, and propionic acid, which are subsequently used by the host as a source of energy [151]. As a result of the fermentation of carbohydrates, Bifidobacterium or Lactobacillus may produce some compounds inhibiting the development of gastrointestinal pathogens, as well as cause a reduction in the intestinal pH [152]. Moreover, Bifidobacterium genus bacteria demonstrate tolerance to the produced SCFAs and reduced pH. Therefore, due to their favourable effect on the development of beneficial intestinal bacteria, the administration of prebiotics may participate in the inhibition of the development of pathogens. There are very few documented study results regarding the inhibition of the development of pathogens by prebiotics. In 1997 and 2003, Bovee-Oudenhoven et al. studied the use of lactulose in the prevention of Salmonella Enteritidis infections on a rat model. Their results indicated that the acidification of the intestine occurring as a result of lactulose fermentation caused the reduced development of pathogens and increased translocation of pathogens from the bowel [153]. It was also demonstrated that the administration of prebiotics increases the absorption of minerals, mostly of and calcium [154,155].

3.4. Prebiotics for Humans The presence of prebiotics in the diet may lead to numerous health benefits. Studies on colorectal carcinoma demonstrated that the disease occurs less commonly in people who often eat vegetables and fruit. This effect is attributed mostly to inulin and oligofructose [156]. Among the advantages of those prebiotics, one may also mention the reduction of the blood LDL (low-density lipoprotein) level, stimulation of the immunological system, increased absorbability of calcium, maintenance of correct intestinal pH value, low caloric value, and alleviation of symptoms of peptic ulcers and vaginal mycosis [157]. Other effects of inulin and oligofructose on human health are: the prevention of carcinogenesis, as well as the support of lactose intolerance or dental caries treatment [131]. Rat studies demonstrated that administration of inulin for five weeks caused a significant reduction of blood triacylglycerol levels [156]. Human studies demonstrated that the daily use of 12 g of inulin for one Nutrients 2017, 9, 1021 14 of 30 month led to the reduction of blood VLDL (very low-density lipoprotein) levels (the reduction of triacylglycerols by 27%, and of cholesterol by 5%). This effect is associated with the effect of the prebiotic on hepatic metabolism and the inhibition of acetyl-CoA carboxylase and of glukose-6-phosphate dehydrogenase. It is also supposed that oligofructose accelerates lipid catabolism [157]. Asahara et al. (2001) demonstrated a protective effect of galactooligosaccharides (GOS) in the prevention of Salmonella Typhimurium infections in a murine model [158]. Buddington et al. (2002) confirmed a positive effect of fructooligosaccharides (FOS) on protection against Salmonella Typhimurium and Listeria monocytogenes infections [159]. Moreover, prebiotics are helpful in combating pathogenic microorganisms, such as Salmonella Enteritidis and Escherichia coli, and reduce odour compounds [160]. There are many reports regarding the positive effect of prebiotics on the carcinogenesis process. Results of rat studies proved that a prebiotic-enriched diet leads to significantly reduced indexes of carcinogenesis. Scientific research demonstrated that butyric acid may be a chemopreventive factor in carcinogenesis [161], or an agent protecting against the development of colorectal carcinoma through the promotion of cell differentiation [162]. Besides butyric acid, propionic acid also may possess anti-inflammatory properties in relation to colorectal carcinoma cells. In vitro studies on human L97 and HT29 cell lines (representing early and late stages of colorectal carcinoma) demonstrated that inulin fractions in plasma supernatant caused a significant inhibition of growth and induction of apoptosis in human colorectal carcinoma [163]. According to scientific reports, the administration of inulin and oligofructose to rats caused the inhibition of azoxymethane-induced colorectal carcinoma at the growth stage [164]. The supplementation of inulin and oligofructose at the dose of 5%–15% had also an effect on reduced occurrence of breast cancer in rats and of metastases to lungs [165]. However, those results have to be confirmed in humans. Table5 lists the results of studies focusing on the effect of prebiotics on human health. There are examples of clinical trials during which the prebiotics group received the prebiotic prophylactically or in addition to the standard therapy.

Table 5. Examples of clinical trials regarding the effect of prebiotics on human health.

Time of References Subjects Prebiotic Main Outcome Administration Obesity There was a reduction in body weight of 1.03 ± 0.43 kg with oligofructose supplementation, whereas the control group experienced an increase in body weight of 0.45 ± 0.31 kg over 48 healthy adults 12 weeks (p = 0.01). Glucose decreased in the oligofructose with a body mass group and increased in the control group between the initial [166] OFS 12 weeks index (in kg/m2) and final tests (p ≤ 0.05). Insulin concentrations mirrored this >25 pattern (p ≤ 0.05). Oligofructose supplementation did not affect plasma active glucagon-like peptide 1 secretion. According to a visual analogue scale designed to assess side effects, oligofructose was well tolerated. Insulin resistance syndrome The plasma glucose response to a fixed exogenous insulin 10 patients with 4 weeks (double [167] FOS bolus did not differ at the end of the two periods. FOS had no type 2 diabetes repetition) effect on glucose and lipid metabolism in type 2 diabetics. Type 2 diabetes 15 subjects with 5 weeks (double A supplement of 15 g/day of AX-rich fibre can significantly [168] AX type 2 diabetes repetition) improve glycaemic control in people with type 2 diabetes. No effects of arabinoxylan were observed for insulin, adiponectin, leptin, or resistin as well as for apolipoprotein B, 11 patients with and unesterified fatty acids. In conclusion, the consumption [169] impaired glucose AX 6 weeks of AX in subjects with impaired glucose tolerance improved tolerance fasting serum glucose and triglycerides. However, this beneficial effect was not accompanied by changes in fasting adipokine concentrations. Non-alcoholic fatty liver disease Compared to placebo, OFS significantly decreased serum 7 patients with aminotransferases, aspartate aminotransferase after 8 weeks, [170] non-alcoholic OFS 8 weeks and insulin level after 4 weeks, but this could not be related to steatohepatitis a significant effect on plasma lipids. Nutrients 2017, 9, 1021 15 of 30

Table 5. Cont.

Time of References Subjects Prebiotic Main Outcome Administration Irritable bowel syndrome (IBS), gastrointestinal disorders, elimination of Helicobacter, inflammatory bowel disease (IBD), diarrhoeas 281 healthy infants Fewer episodes of acute diarrhoea, fewer upper respiratory [171] GOS, FOS 12 months (15 to 120 days) tract infections. 160 healthy Prebiotic formula well tolerated, normal growth trend toward bottle-fed infants [172] GOS, FOS 3 months a higher percentage of Bifidobacterium and a lower percentage within 0–14 days of E. coli in stool, suppresses Clostridium in stool. after birth The concentration of secretory IgA was higher in the prebiotic [173] 215 healthy infants GOS, FOS 27 weeks group than the control; also, Bifidobacterium percentage was higher than the control and Clostridium was lower. Inulin treatment resulted in decreased endoscopic and 24 patients with histological inflammation. This effect was associated with [174] inulin 3 weeks chronic pouchitis increased intestinal butyrate, lowered pH, and significantly decreased numbers of Bacteroides fragilis. 10 Crohn’s [175] FOS 3 weeks Reduced disease activity index. disease patients Atopic dermatitis 259 infants at risk [176] GOS, FOS 6 months Significant reduction of frequency of AD. for atopy 259 healthy term Prebiotic group had significantly lower allergic infants with a [177] GOS, FOS 6 months symptoms—AD, wheezing, urticaria, and fewer upper parental history respiratory infections than controls during the first 2 years. of atopy Alleviation of lactose intolerance 71% of subjects reported improvements in at least one symptom (pain, , diarrhoea, cramping, or flatulence). 85 lactose Also on day 36, populations of bifidobacteria significantly [178] intolerant GOS 36 days increased by 90% in 27 of the 30 non-lactose tolerant participants participants who took GOS. Lactose fermenting Bifidobacterium, Faecalibacterium, and Lactobacillus were all significantly increased. Different types of cancer and side effects associated with cancer Human L97 and HT29 cell lines (representing early Growth inhibition and induction of apoptosis in human [163] inulin No data and late stages of colorectal carcinoma. colorectal carcinoma) Abbreviations: AD—atopic dermatitis; AX—arabinoxylan; FOS—fructooligosaccharides; GOS—galactooligosaccharides; IgA—immunoglobulin A; OFS—oligofructose.

4. Synbiotics Synbiotics are used not only for the improved survival of beneficial microorganisms added to food or feed, but also for the stimulation of the proliferation of specific native bacterial strains present in the gastrointestinal tract [179]. The effect of synbiotics on metabolic health remains unclear. It should be mentioned that the health effect of synbiotics is probably associated with the individual combination of a probiotic and prebiotic [180]. Considering a huge number of possible combinations, the application of synbiotics for the modulation of intestinal microbiota in humans seems promising [181].

4.1. Synbiotic Selection Criteria The first aspect to be taken into account when composing a synbiotic formula should be a selection of an appropriate probiotic and prebiotic, exerting a positive effect on the host’s health when used separately. The determination of specific properties to be possessed by a prebiotic to have a favourable effect on the probiotic seems to be the most appropriate approach. A prebiotic should selectively stimulate the growth of microorganisms, having a beneficial effect on health, with simultaneous absent (or limited) stimulation of other microorganisms. Nutrients 2017, 9, 1021 16 of 30

4.2. Synbiotics in Use Previous sections discussed probiotic microorganisms and prebiotic substances most commonly used in human nutrition. A combination of Bifidobacterium or Lactobacillus genus bacteria with fructooligosaccharides in synbiotic products seems to be the most popular. Table4 presents the most commonly used combinations of probiotics and prebiotics.

4.3. Mechanism of Action of Synbiotics Considering the fact that a probiotic is essentially active in the small and , and the effect of a prebiotic is observed mainly in the large intestine, the combination of the two may have a synergistic effect [182]. Prebiotics are used mostly as a selective medium for the growth of a probiotic strain, fermentation, and intestinal passage. There are indications in the literature that, due to the use of prebiotics, probiotic microorganisms acquire higher tolerance to environmental conditions, including: oxygenation, pH, and temperature in the intestine of a particular organism [183]. However, the mechanism of action of an extra energy source that provides higher tolerance to these factors is not sufficiently explained. That combination of components leads to the creation of viable microbiological dietary supplements, and ensuring an appropriate environment allows a positive impact on the host’s health. Two modes of synbiotic action are known [184]:

(1) Action through the improved viability of probiotic microorganisms; (2) Action through the provision of specific health effects.

The stimulation of probiotics with prebiotics results in the modulation of the metabolic activity in the intestine with the maintenance of the intestinal biostructure, development of beneficial microbiota, and inhibition of potential pathogens present in the gastrointestinal tract [180]. Synbiotics result in reduced concentrations of undesirable metabolites, as well as the inactivation of nitrosamines and cancerogenic substances. Their use leads to a significant increase of levels of short-chain fatty acids, ketones, carbon disulphides, and methyl acetates, which potentially results in a positive effect on the host’s health [184]. As for their therapeutic efficacy, the desirable properties of synbiotics include antibacterial, anticancerogenic, and anti-allergic effects. They also counteract decay processes in the intestine and prevent constipation and diarrhoea. It turns out that synbiotics may be highly efficient in the prevention of osteoporosis, reduction of blood fat and sugar levels, regulation of the immunological system, and treatment of brain disorders associated with abnormal hepatic function [185]. The concept of mechanisms of synbiotic action, based on the modification of intestinal microbiota with probiotic microorganisms and appropriately selected prebiotics as their substrates, is presented in Figure1.

4.4. Synbiotics for Humans Synbiotics have the following beneficial effects on humans [186]:

(1) Increased Lactobacillus and Bifidobacterium genus count and maintenance of balance of the intestinal microbiota; (2) Improved hepatic function in patients suffering from cirrhosis; (3) Improved immunomodulative abilities; (4) Prevention of bacterial translocation and reduced incidence of nosocomial infections in patients’ post-surgical procedures and similar interventions.

The translocation of bacterial metabolism products, such as lipopolysaccharides (LPSs), ethanol, and short-chain fatty acids (SFCAs), leads to their penetration of the liver. SCFAs also stimulate the synthesis and storage of hepatic triacylglycerols. Those processes may intensify the mechanisms of hepatic detoxication, which may result in hepatic storage of triacylglycerol (IHTG), and intensify steatosis of the organ. A randomised trial on the use of a synbiotic containing five probiotics (Lactobacillus plantarum, Lactobacillus delbrueckii spp. bulgaricus, Lactobacillus acidophilus, Nutrients 2017, 9, 1021 17 of 30

Lactobacillus rhamnosus, Bifidobacterium bifidum) and inulin as a prebiotic in adult subjects with NASH (non-alcoholic steatohepatisis) demonstrated a significant reduction of IHTG (intrahepatic triacylglycerol) within six months [187]. It is also known that LPSs induce proinflammatory cytokines, such as the tumour necrosis factor alpha (TNF-α), playing a crucial role in insulin resistance and inflammatory cell uptake in NAFLD (non-alcoholic fatty liver disease). In the study on the effect of the synbiotic product containing a blend of probiotics (, Lactobacillus rhamnosus, Streptococcus thermophilus, Bifidobacterium breve, Lactobacillus acidophilus, Bifidobacterium longum, Lactobacillus bulgaricus) and fructooligosccharides, 52 adults participated for 28 weeks. It was found that supplementation with the synbiotic resulted in the inhibition of NF-κB (nuclear factor κB) and reduced production of TNF-α (tumour necrosis factor α)[188]. In rat studies, an increased level of intestinal IgA was found, following the introduction of the synbiotic product containing Lactobacillus rhamnosus and Bifidobacterium lactis, and inulin and oligofructose as prebiotics to the diet. Synbiotics lead to reduced blood cholesterol levels and lower blood pressure [157]. Moreover, synbiotics are used in the treatment of hepatic conditions [189] and improve the absorption of calcium, magnesium, and phosphorus [190]. Danq et al. (2013), in a meta-analysis, evaluated published studies on pro/prebiotics for eczema prevention, investigating bacterial strain efficacy and changes to the allergy status of the children involved. This meta-analysis found that probiotics or synbiotics may reduce the incidence of eczema in infants aged <2 years. Systemic sensitization did not change following probiotic administration [191]. Studies carried out within the framework of the SYNCAN project funded by the European Union verified the anti-carcinogenic properties of synbiotics. The effect of fructooligosaccharides (SYN1) combined with two probiotic strains (Lactobacillus rhamnosus GG and Bifidobacterium animalis subsp. lactis Bb12) on the health of patients at risk of colorectal cancer was studied. As a result, a change of biomarkers (genotoxicity, labelling index, labelled cells/crypt, transepithelial resistance, necrosis, interleukin 2, interferon γ) indicating the development of the disease in cancer patients, and in patients post polyp excision, was observed [192]. It was concluded that the application of the studied synbiotic may reduce the risk of colorectal carcinoma. A lower level of DNA damage was also observed, as well as a lower colonocyte proliferation ratio [147]. Table6 lists the results of studies focusing on the effect of synbiotics on human health. There are examples of clinical trials during which the synbiotics group received the synbiotic prophylactically or in addition to the standard therapy.

Table 6. Examples of clinical trials regarding the effect of synbiotics on human health.

Time of References Subjects Composition of Synbiotic Main Outcome Administration Obesity 153 obese men L. rhamnosus Weight loss and reduction in leptin. [193] 36 weeks and women CGMCC1.3724, inulin Increase in Lachnospiraceae. L. casei, L. rhamnosus, 70 children and S. thermophilus, B. breve, Decrease in BMI z-score and [194] adolescents with 8 weeks L. acidophilus, B. longum, waist circumference. high BMI L. bulgaricus, FOS Changes in anthropometric L. acidophilus, L. rhamnosus, measurements. Decrease in TC, [195] 77 obese children B. bifidum, B. longum, 4 weeks LDL-C, and total oxidative stress E. faecium, FOS serum levels. Insulin resistance syndrome L. casei, L. rhamnosus, The levels of fasting blood sugar and S. thermophilus, B. breve, [196] 38 subjects with IRS 28 weeks insulin resistance L. acidophilus, B. longum, improved significantly. L. bulgaricus, FOS Nutrients 2017, 9, 1021 18 of 30

Table 6. Cont.

Time of References Subjects Composition of Synbiotic Main Outcome Administration Type 2 diabetes L. acidophilus, L. casei, L. rhamnosus, L. bulgaricus, Increased HOMA-IR and TGL plasma [197] 54 patients with T2D 8 weeks B. breve, B. longum, level; reduced CRP in serum. S. thermophilus, FOS Significant reduction in serum insulin [198] 81 patients with T2D L. sporogenes, inulin 8 weeks levels, HOMA-IR, and homeostatic model assessment cell function. Decrease in serum lipid profile (TAG, [199] 78 patients with T2D L. sporogenes, inulin 8 weeks TC/HDL-C) and a significant increase in serum HDL-C levels. L. acidophilus, B. bifidum, Increased HDL-C and reduced [200] 20 patients with T2D 2 weeks oligofructose fasting glycaemia. Non-alcoholic fatty liver disease L. plantarum, L. delbrueckii spp. 20 individuals [187] bulgaricus, L. acidophilus, 26 weeks Decreased IHTG content. with NASH L. rhamnosus, B. bifidum, inulin L. casei, L. rhamnosus, 52 adult individuals S. thermophilus, B. breve, Inhibition of NF-κB and reduction [188] 30 weeks with NAFLD L. acidophilus, B. longum, of TNF-α. L. bulgaricus, FOS Irritable bowel syndrome (IBS), gastrointestinal disorders, elimination of Helicobacter, inflammatory bowel disease (IBD), diarrhoeas On average, an 18% improvement in total IBS-QoL score was reported and significant improvements in bloating severity, satisfaction with bowel L. acidophilus La-5®, B. animalis movements, and the severity of IBS [201] 76 patients with IBS ssp. lactis BB-12®, dietary 4 weeks symptoms’ interference with patients’ fibres (Beneo) everyday life were observed. However, there were no statistically significant differences between the synbiotic group and the placebo group. From a total of 69 H. pylori-infected children (female/male = 36/33; mean ± SD = 11.2 ± 3.0 years), eradication 69 children aged 6–16 was achieved in 20 out of 34 years who had biopsy participants in the standard therapy [202] B. lactis B94, inulin 14 days proven H. pylori group and 27/35 participants in the infection synbiotic group. There were no significant differences in eradication rates between the standard therapy and the synbiotic groups. Patients with UC on synbiotic therapy experienced greater quality-of-life [203] 40 patients with UC B. longum, psyllium 4 weeks changes than patients on probiotic or prebiotic treatment. Atopic dermatitis This synbiotic mixture did not have a B. breve M-16V, GOS and FOS beneficial effect on AD severity in [204] 90 infants with AD 12 weeks mixture (Immunofortis®) infants, although it did successfully modulate their intestinal microbiota. L. casei, L. rhamnosus, S. 40 infants and children A mixture of seven probiotic strains thermophilus, B. breve, [205] aged 3 months to 8 weeks and FOS may clinically improve the L. acidophilus, B. infantis, 6 years with AD severity of AD in young children. L. bulgaricus, FOS Nutrients 2017, 9, 1021 19 of 30

Table 6. Cont.

Time of References Subjects Composition of Synbiotic Main Outcome Administration Alleviation of lactose intolerance Consumption of the probiotic mixture improved the gastrointestinal performance associated with lactose load in subjects with LI. Symptoms Lactobacillus, [206] 20 females and males 5 weeks were additionally reduced by the Bifidobacterium, FOS addition of prebiotics. The supplementation was safe and well tolerated, with no significant adverse effect observed. Different types of cancer and side effects associated with cancer Increased L. rhamnosus and B. lactis in faeces, reduction in C. perfringens, 43 polypeptomized L. rhamnosus GG, B. lactis prevents increased secretion of IL-2 in [192] and 37 colon 12 weeks Bb12, inulin polypectomized patients, increased cancer patients production of interferon-γ in cancer patients. Abbreviations: BMI—body mass index; CFU—colony-forming-unit; CRP—C-reactive protein; FOS—fructo-oligossacharides; IBS-QoL—quality of life with IBS; HDL-C—high-density lipoprotein cholesterol; HOMA-IR—homeostasis model assessment of insulin resistance; IHTG—intrahepatic triacylglycerol; IRS—insulin resistance syndrome; LDL-C—low-density lipoprotein cholesterol; LI—lactose intolerance; NAFLD—non-alcoholic fatty liver disease; NF-κB—nuclear factor κB; T2D—type 2 diabetes; TAG—triacylglycerols; TC—total cholesterol; TGL—total glutathione levels; TNF-α—tumour necrosis factor α; UC—ulcerative colitis.

5. Summary Probiotic organisms are crucial for the maintenance of balance of human intestinal microbiota. Numerous scientific reports confirm their positive effect in the host’s health. Probiotic microorganisms are attributed a high therapeutic potential in, e.g., obesity, insulin resistance syndrome, type 2 diabetes, and non-alcohol hepatic steatosis [207]. It seems also that probiotics may be helpful in the treatment of irritable bowel syndrome, enteritis, bacterial infections, and various gastrointestinal disorders and diarrhoeas. Probiotic microorganisms are also effective in the alleviation of lactose intolerance and the treatment of atopic dermatitis. A positive effect of probiotics in the course of various neoplastic diseases and side effects associated with anti-cancer therapies is also worth noting. Prebiotics may be used as an alternative to probiotics, or as an additional support for them. It turns out that the development of bio-therapeutic formulas containing both appropriate microbial strains and synergistic prebiotics may lead to the enhancement of the probiotic effect in the small intestine and the colon. Those “enhanced” probiotic products may be even more effective, and their protective and stimulatory effect superior to their components administered separately [208]. It seems that we will see further studies on combinations of probiotics and prebiotics, and further development of synbiotics. Future studies may explain the mechanisms of actions of those components, which may confer a beneficial effect on human health.

Acknowledgments: We would like to thank the National Centre for Research and Development for the financial support of publication of this paper within the project PBS3/A8/32/2015 realized within the framework of the Program of Applied Studies. Conflicts of Interest: The authors declare no conflict of interest.

References

1. Vergin, F. Anti-und Probiotica. Hipokrates 1954, 25, 116–119. 2. Lilly, D.M.; Stillwell, R.H. Probiotics: Growth promoting factors produced by microorganisms. Science 1965, 147, 747–748. [CrossRef][PubMed] 3. Fuller, R. Probiotics in man and animals. J. Appl. Microbiol. 1989, 66, 365–378. 4. Guarner, F.; Schaafsma, G.J. Probiotics. Int. J. Food Microbiol. 1998, 39, 237–238. [CrossRef] Nutrients 2017, 9, 1021 20 of 30

5. Food and Agriculture Organization (FAO). Guidelines for the Evaluation of Probiotics in Food; Report of a Joint FAO/WHO Working Group on Drafting Guidelines for the Evaluation of Probiotics in Food; FAO: London, ON, Canada, 30 April–1 May 2002. 6. Hill, C.; Guarner, F.; Reid, G.; Gibson, G.R.; Merenstein, D.J.; Pot, B.; Morelli, L.; Canani, R.B.; Flint, H.J.; Salminen, S.; et al. Expert consensus document: The International Scientific Association for Probiotics and Prebiotics consensus statement on the scope and appropriate use of the term probiotic. Nat. Rev. Gastroenterol. Hepatol. 2014, 11, 506–514. [CrossRef][PubMed] 7. Gibson, R.G.; Roberfroid, M.B. Dietary modulation of the human colonic microbiota: Introducing the concept of prebiotics. J. Nutr. 1995, 125, 1401–1412. [PubMed] 8. Gibson, G.R.; Probert, H.M.; van Loo, J.; Rastall, R.A.; Roberfroid, M. Dietary modulation of the human colonic microbiota: Updating the concept of the prebiotics. Nutr. Res. Rev. 2004, 17, 259–275. [CrossRef] [PubMed] 9. Food and Agriculture Organization. FAO Technical Meeting on Prebiotics: Food Quality and Standards Service (AGNS), Food and Agriculture Organization of the United Nations (FAO); FAO Technical Meeting Report; FAO: Rome, Italy, 15–16 September 2007. 10. Skalkam, M.L.; Wiese, M.; Nielsen, D.S.; van Zanten, G. In Vitro Screening and Evaluation of Synbiotics; University of Copenhagen: Copenhagen, Denmark, 2016; Chapter 33; pp. 477–486. 11. Cencic, A.; Chingwaru, W. The role of functional foods, , and food supplements in intestinal health. Nutrients 2010, 2, 611–625. [CrossRef][PubMed] 12. Rioux, K.P.; Madsen, K.L.; Fedorak, R.N. The role of enteric microflora in inflammatory bowel disease: Human and animal studies with probiotics and prebiotics. Gastroenterol. Clin. N. Am. 2005, 34, 465–482. [CrossRef][PubMed] 13. Bengmark, S. Bioecological control of the gastrointestinal tract: The role of flora and supplemented probiotics and synbiotics. Gastroenterol. Clin. N. Am. 2005, 34, 413–436. [CrossRef][PubMed] 14. Panesar, P.S.; Kaur, G.; Panesar, R.; Bera, M.B. Synbiotics: Potential Dietary Supplements in Functional Foods; IFIS: Berkshire, UK, 2009. [CrossRef] 15. Hosono, A. Fermented milk in the orient. In Functions of Fermented Milk: Challengers for the Health Sciences; Nakazawa, Y., Hosono, A., Eds.; Elsevier Science Publishers Ltd.: Barking, UK, 1992; pp. 61–78. 16. Miecznikow, E. O naturze ludzkiej—Zarys Filozofii Optymistycznej; Translation F. Wermi´nski;Wydawnictwo Biblioteka Naukowa: Warszawa, Poland, 1907. 17. Lee, Y.K. Selection and maintenance of probiotic microorganisms. In Handbook of Probiotics and Prebiotics; Lee, Y.K., Salminen, S., Eds.; Wiley-VCH: Weinheim, Germany, 2009; pp. 177–187. [CrossRef] 18. Sanders, M.E.; Gibson, G.; Harsharnjit, S.G.; Guarner, F. Probiotics: Their Potential to Impact Human Health; CAST Issue Paper; CAST: Ames, IA, USA, 2007; Volume 36, pp. 1–20. 19. Sanders, M.E.; Lenoir-Wijnkoop, I.; Salminen, S.; Merenstein, D.J.; Gibson, G.R.; Petschow, B.W.; Nieuwdorp, M.; Tancredi, D.J.; Cifelli, C.J.; Jacques, P.; et al. Probiotics and prebiotics: Prospects for public health and nutritional recommendations. Ann. N. Y. Acad. Sci. 2014, 1309, 19–29. [CrossRef][PubMed] 20. European Food Safety Authority (EFSA). Opinion of the Scientific Committee on a request from EFSA related to a generic approach to the safety assessment by EFSA of microorganisms used in food/feed and the production of food/feed additives. EFSA J. 2005, 226, 1–12. 21. Simon, O. Micro-organisms as feed additives—Probiotics. Adv. Pork Prod. 2005, 16, 161–167. 22. Anadón, A.; Martínez-Larra´naga,M.R.; Martínez, M.A. Probiotics for animal nutrition in the European Union. Regulation and safety assessment. Regul. Toxicol. Pharmacol. 2006, 45, 91–95. [CrossRef][PubMed] 23. Gaggia, F.; Mattarelli, P.; Biavati, B. Probiotics and prebiotics in animal feeding for safe food production. Int. J. Food Microbiol. 2010, 141, S15–S28. [CrossRef][PubMed] 24. European Food Safety Authority (EFSA). The European Union Summary report on trends and sources of zoonoses, zoonotic agents and food-borne outbreaks in 2011. EFSA J. 2013, 3129, 1–250. 25. European Food Safety Authority (EFSA). Scientific opinion on the maintenance of the list of QPS biological agents intentionally added to food and feed (2013 update). EFSA J. 2013, 15, 1–108. 26. European Food Safety Authority (EFSA). Scientific Opinion on the update of the list of QPS-recommended biological agents intentionally added to food or feed as notified to EFSA (2017 update). EFSA J. 2017, 15, 1–177. Nutrients 2017, 9, 1021 21 of 30

27. Schachtsiek, M.; Hammes, W.P.; Hertel, C. Characterization of Lactobacillus coryniformis DSM 20001T surface protein CPF mediating coaggregation with and aggregation among pathogens. Appl. Environ. Microbiol. 2004, 70, 7078–7085. [CrossRef][PubMed] 28. Oelschlaeger, T.A. Mechanisms of probiotic actions—A review. Int. J. Med. Microbiol. 2010, 300, 57–62. [CrossRef][PubMed] 29. Cremonini, F.; di Caro, S.; Nista, E.C.; Bartolozzi, F.; Capelli, G.; Gasbarrini, G.; Gasbarrini, A. Meta-analysis: The eqect of probiotic administration on antibiotic associateddiarrhoea. Aliment. Pharmacol. Ther. 2002, 16, 1461–1467. [CrossRef][PubMed] 30. Johnston, B.C.; Supina, A.L.; Vohra, S. Probiotics for pediatric antibiotic-associated diarrhea: A meta-analysis of randomized placebo-controlled trials. Can. Med. Assoc. J. 2006, 175, 377–383. [CrossRef][PubMed] 31. Schoster, A.; Kokotovic, B.; Permin, A.; Pedersen, P.D.; Dal Bello, F.; Guardabassi, L. In Vitro inhibition of Clostridium difficile and Clostridium perfringens by commercial probiotic strains. Anaerobe 2013, 20, 36–41. [CrossRef][PubMed] 32. JimmySaint-Cyr, M.; Haddad, N.; Taminiau, B.; Poezevara, T.; Quesne, S.; Amelot, M.; Daube, G.; Chemaly, M.; Dousset, X.; Guyard-Nicodème, M. Use of the potential probiotic strain Lactobacillus salivarius SMXD51 to control Campylobacter jejuni in broilers. Int. J. Food Microbiol. 2017, 247, 9–17. [CrossRef][PubMed] 33. Carter, A.; Adams, M.; La Ragione, R.M.; WoodWard, M.J. Colonisation of poultry by Salmonella Enteritidis S1400 is reduced by combined administration of Lactobacillus salivarius 59 and Enterococcus faecium PXN-33. Vet. Microbiol. 2017, 199, 100–107. [CrossRef][PubMed] 34. Chingwaru, W.; Vidmar, J. Potential of Zimbabwean commercial probiotic products and strains of Lactobacillus plantarum as prophylaxis and therapy against diarrhoea caused by Escherichia coli in children. Asian Pac. J. Trop. Med. 2017, 10, 57–63. [CrossRef][PubMed] 35. Hussain, S.A.; Patil, G.R.; Reddi, S.; Yadav, V.; Pothuraju, R.; Singh, R.R.B.; Kapila, S. Aloe vera (Aloe barbadensis Miller) supplemented probiotic lassi prevents Shigella infiltration from epithelial barrier into systemic blood flow in mice model. Microb. Pathog. 2017, 102, 143–147. [CrossRef][PubMed] 36. Sikorska, H.; Smoragiewicz, W. Role of probiotics in the prevention and treatment of ethicillin-resistant Staphylococcus aureus infections. Int. J. Antimicrob. Agents 2013, 42, 475–481. [CrossRef][PubMed] 37. De Montijo-Prieto, S.; Moreno, E.; Bergillos-Meca, T.; Lasserrot, A.; Ruiz-López, M.; Ruiz-Bravo, A.; Jimenez-Valera, M. A Lactobacillus plantarum strain isolated from kefir protects against intestinal infection with Yersinia enterocolitica O9 and modulates immunity in mice. Res. Microbiol. 2015, 166, 626–632. [CrossRef] [PubMed] 38. Thomas, D.W.; Greer, F. Probiotics and prebiotics in pediatrics. Pediatrics 2010, 126, 1217–1231. [CrossRef] [PubMed] 39. Heczko, P.; Strus, M.; Jawie´n,M.; Szyma´nski,H. Medyczne zastosowanie probiotyków. Wiad. Lek. 2005, 58, 640–646. [PubMed] 40. Kumar, S.; Bansal, A.; Chakrabarti, A.; Singhi, S. Evaluation of efficacy of probiotics in prevention of Candida colonization in a PICU—A randomized controlled trial. Crit. Care Med. 2013, 41, 565–572. [CrossRef] [PubMed] 41. Nase, L.; Hatakka, K.; Savilahti, E. Eqect of long-term consumption of Lactobacillus GG in milk on dental caries and caries risk in children. Caries Res. 2001, 35, 412–420. [CrossRef][PubMed] 42. Li, P.; Gu, Q. Complete genome sequence of Lactobacillus plantarum LZ95, a potential probiotic strain producing bacteriocins and B-group riboflavin. J. Biotechnol. 2016, 229, 1–2. [CrossRef][PubMed] 43. Gu, Q.; Zhang, C.; Song, D.; Li, P.; Zhu, X. Enhancing vitamin B12 content in soy-yogurt by Lactobacillus reuteri. Int. J. Food Microbiol. 2015, 206, 56–59. [CrossRef][PubMed] 44. Pompei, A.; Cordisco, L.; Amaretti, A.; Zanoni, S.; Matteuzzi, D.; Rossi, M. production by bifidobacteria as a potential probiotic property. Appl. Environ. Microbiol. 2007, 73, 179–185. [CrossRef][PubMed] 45. Nova, E.; Warnberg, J.; Gomez-Martinez, S.; Diaz, L.E.; Romeo, J.; Marcos, A. Immunodulatory effects of probiotics in different stages of life. Br. J. Nutr. 2007, 98, S90–S95. [CrossRef][PubMed] 46. Mishra, C.; Lambert, J. Production of anti-microbial substances by probiotics. Asia Pac. J. Clin. Nutr. 1996, 5, 20–24. [PubMed] 47. Ouwehand, A.C.; Kirjavainen, P.V.; Shortt, C.; Salminen, S. Probiotics: Mechanisms and established effects. Int. Dairy J. 1999, 9, 43–52. [CrossRef] Nutrients 2017, 9, 1021 22 of 30

48. Reid, G.; McGroarty, J.A.; Angotti, R.; Cook, R.L. Lactobacillus inhibitor production against Escherichia coli and coaggregation ability with uropathogens. Can. J. Microbiol. 1988, 34, 344–351. [CrossRef][PubMed] 49. Schellenberg, J.; Smoragiewicz, W.; Karska-Wysocki, B. A rapid method combining immunofluorescence and flow cytometry for improved understanding of competitive interactions between lactic acid bacteria (LAB) and methicillin-resistant S. aureus (MRSA) in mixed culture. J. Microbiol. Methods 2006, 65, 1–9. [CrossRef] [PubMed] 50. Ishikawa, H.; Akedo, I.; Otani, T.; Suzuki, T.; Nakamura, T.; Takeyama, I.; Ishiguro, S.; Miyaoka, E.; Sobue, T.; Kakizoe, T. Randomized trial of dietary fiber and Lactobacillus casei administration for prevention of colorectal tumors. Int. J. Cancer 2005, 116, 762–767. [CrossRef][PubMed] 51. Vandenbergh, P.A. Lactic acid bacteria, their metabolic products and interference with microbial growth. FEMS Microbiol. Rev. 1993, 12, 221–238. [CrossRef] 52. Guillot, J.F. Probiotic feed additives. J. Vet. Pharmacol. Ther. 2003, 26, 52–55. 53. Isolauri, E.; Sutas, Y.; Kankaanpaa, P.; Arvilommi, H.; Salminen, S. Probiotics: Effects on immunity. Am. J. Clin. Nutr. 2001, 73, 444–450. [CrossRef][PubMed] 54. Brandao, R.L.; Castro, I.M.; Bambirra, E.A.; Amaral, S.C.; Fietto, L.G.; Tropia, M.J.M. Intracellular signal triggered by cholera toxin in Saccharomyces boulardii and Saccharomyces cerevisiae. Applied Environ. Microbiol. 1998, 64, 564–568. [PubMed] 55. Collado, M.C.; Meriluoto, J.; Salminen, S. Interactions between pathogens and lactic acid bacteria: Aggregation and coaggregation abilities. Eur. J. Food Res. Technol. 2007, 226, 1065–1073. [CrossRef] 56. Begley, M.; Hill, C.; Gahan, C.G.M. Bile salt hydrolase activity in probiotics. Appl. Environ. Microbiol. 2006, 72, 1729–1738. [CrossRef][PubMed] 57. Weinberg, E.D. The Lactobacillus anomaly: Total iron abstinence. Perspect. Biol. Med. 1997, 40, 578–583. [CrossRef][PubMed] 58. Elli, M.; Zink, R.; Rytz, A.; Reniero, R.; Morelli, L. Iron requirement of Lactobacillus spp. in completely chemically defined growth media. J. Appl. Microbiol. 2000, 88, 695–703. [CrossRef][PubMed] 59. Borchers, A.T.; Selmi, C.; Meyers, F.J.; Keen, C.L.; Gershwin, M.E. Probiotics and immunity. J. Gastroenterol. 2009, 44, 26–46. [CrossRef][PubMed] 60. Gill, H.S.; Cross, M.L. Probiotics and immune function. In Nutrition and Immune Function; Calder, P.C., Field, C.J., Gill, H.S., Eds.; CABI Publishing: Wallingford, UK, 2002; pp. 251–272. 61. Marteau, P.; Shanahan, F. Basic aspects and pharmacology of probiotics: An overview of pharmacokinetics, mechanisms of action and side-effects. Best Pract. Res. Clin. Gastroenterol. 2003, 17, 725–740. [CrossRef] 62. Schatzmayr, G.; Zehner, F.; Taubel, M.; Schatzmayr, D.; Klimitsch, A.; Loibner, A.P.; Binder, E.M. Microbiologicals for deactivating mycotoxins. Mol. Nutr. Food Res. 2006, 50, 543–551. [CrossRef][PubMed] 63. McCormick, S.P. Microbial detoxification of mycotoxins. J. Chem. Ecol. 2013, 39, 907–918. [CrossRef] [PubMed] 64. Nikbakht Nasrabadi, E.; Jamaluddin, R.; Abdul Mutalib, M.S.; Khaza’ai, H.; Khalesi, S.; Mohd Redzwan, S. Reduction of aflatoxin level in aflatoxin-induced rats by the activity of probiotic Lactobacillus casei strain Shirota. J. Appl. Microbiol. 2013, 114, 1507–1515. [CrossRef][PubMed] 65. Upadrasta, A.; Madempudi, R.S. Probiotics and blood pressure: Current insights. Integr. Blood Press. Control 2016, 9, 33–42. [CrossRef][PubMed] 66. Khalesi, S.; Sun, J.; Buys, N.; Jayasinghe, R. Effect of probiotics on blood pressure. A Systematic review and meta-analysis of randomized, controlled trials. Hypertension 2014, 70, 1–13. [CrossRef][PubMed] 67. Ruan, Y.; Sun, J.; He, J.; Chen, F.; Chen, R.; Chen, H. Effect of probiotics on glycemic control: A systematic review and meta-analysis of randomized, controlled trials. PLoS ONE 2015, 10, e0132121. [CrossRef] [PubMed] 68. Lima-Filho, J.V.M.; Vieira, E.C.; Nicoli, J.R. Saccharomyces boulardii and Escherichia coli combinations against experimential infections with Shigella flexneri and Salmonella enteritidis subsp. Typhimurium. J. Appl. Microbiol. 2000, 88, 365–370. [CrossRef] 69. Bengmark, S. Bioecological control of inflammatory bowel disease. Clin. Nutr. 2007, 26, 169–181. [CrossRef] [PubMed] 70. Geier, M.S.; Butler, R.N.; Howarth, G.S. Inflammatory bowel disease: Current insights into pathogenesis and new therapeutic options; probiotics, prebiotics and synbiotics. Int. J. Food Microbiol. 2007, 115, 1–11. [CrossRef][PubMed] Nutrients 2017, 9, 1021 23 of 30

71. Montalto, M.; Curigliano, V.; Santoro, L.; Vastola, M.; Cammarota, G.; Manna, R.; Gasbarrini, A.; Gasbarrini, G. Management and treatment of lactose malabsorption. World J. Gastroenterol. 2006, 12, 187. [CrossRef] [PubMed] 72. Levri, K.M.; Ketvertis, K.; Deramo, M.; Merenstein, J.H.; D’Amico, F. Do probiotics reduce adult lactose intolerance? J. Family Pract. 2005, 54, 613–620. 73. Geier, M.S.; Butler, R.N.; Howarth, G.S. Probiotics, prebiotics and synbiotics: A role in chemoprevention for colorectal cancer? Cancer Biol. Ther. 2006, 5, 1265–1269. [CrossRef][PubMed] 74. Lesbros-Pantoflickova, D.; Corthesy-Theulaz, I.; Blum, A.L. Helicobacter pylori and probiotics. J. Nutr. 2007, 137, 812S–818S. [PubMed] 75. Guarner, F.; Khan, A.G.; Garisch, J.; Eliakim, R.; Gangl, A.; Thomson, A.; Krabshuis, J.; Lemair, T. Probiotics and Prebiotics; World Gastroenterology Organisation Global Guidelines; World Gastroenterology Organisation: Milwaukee, WI, USA, 2011; Volume 46, pp. 468–481. 76. Falagas, M.E.; Betsi, G.I.; Tokas, T.; Athanasiou, S. Probiotics for prevention of recurrent urinary tract infections in women: A review of the evidence from microbiological and clinical studies. Drugs 2006, 66, 1253–1261. [CrossRef][PubMed] 77. Anukam, K.C.; Osazuwa, E.; Osemene, G.I.; Ehigiagbe, F.; Bruce, A.W.; Reid, G. Clinical study comparing probiotic Lactobacillus GR-1 and RC-14 with metronidazole vaginal gel to treat symptomatic bacterial vaginosis. Microbes Infect. 2006, 8, 2772–2776. [CrossRef][PubMed] 78. Reid, G.; Bruce, A.W. Probiotics to prevent urinary tract infections: The rationale and evidence. World J. Urol. 2005, 24, 28–32. [CrossRef][PubMed] 79. Kukkonen, K.; Savilahti, E.; Haahtela, T.; Juntunen-Backman, K.; Korpela, R.; Poussa, T.; Tuure, T.; Kuitunen, M. Probiotics and prebiotic galacto-oligosaccharides in the prevention of allergic diseases: A randomized, double-blind, placebo-controlled trial. J. Allergy Clin. Immunol. 2007, 119, 192–198. [CrossRef] [PubMed] 80. Simons, L.A.; Amansec, S.G.; Conway, P. Effect of Lactobacillus fermentum on serum lipids in subjects with elevated serum cholesterol. Nutr. Metab. Cardiovasc. Dis. 2006, 16, 531–535. [CrossRef][PubMed] 81. Lieske, J.C.; Goldfarb, D.S.; de Simone, C.; Regnier, C. Use of a probiotic to decrease enteric hyperoxaluria. Kidney Int. 2005, 68, 1244–1249. [CrossRef][PubMed] 82. Rousseaux, C.; Thuru, X.; Gelot, A.; Barnich, N.; Neut, C.; Dubuquoy, L.; Dubuquoy, C.; Merour, E.; Geboes, K.; Chamaillard, M.; et al. Lactobacillus acidophilus modulates intestinal pain and induces opioid and cannabinoid receptors. Nat. Med. 2006, 13, 35–37. [CrossRef][PubMed] 83. Billoo, A.G.; Memon, M.A.; Khaskheli, S.A.; Murtaza, G.; Iqbal, K.; Saeed Shekhani, M.; Siddiqi, A.Q. Role of a probiotic (Saccharomyces boulardii) in management and prevention of diarrhoea. World J. Gastroenterol. 2006, 12, 4557–4560. [CrossRef][PubMed] 84. Parvez, S.; Malik, K.A.; Ah Kang, S.; Kim, H.Y. Probiotics and their fermented food products are beneficial for health. J. Appl. Microbiol. 2006, 100, 1171–1185. [CrossRef][PubMed] 85. McFarland, L.V. Meta-Analysis of probiotics for the prevention of antibiotic associated diarrhea and the treatment of Clostridium difficile disease. Am. J. Gastroenterol. 2006, 101, 812–822. [CrossRef][PubMed] 86. Hatakka, K.; Savilahti, E.; Pönkä, A.; Meurman, J.H.; Poussa, T.; Näse, L.; Saxelin, M.; Ko, R. Effect of long term consumption of probiotic milk on infections in children attending day care centres: Double blind, randomised trial. Br. Med. J. 2001, 322, 1327. [CrossRef] 87. Olivares, M.; Diaz-Ropero, M.; Gomez, N.; Sierra, S.; Lara-Villoslada, F.; Martin, R.; Miguel Rodriguez, J.; Xaus, J. Dietary deprivation of fermented foods causes a fall in innate immune response. Lactic acid bacteria can counteract the immunological effect of this deprivation. J. Dairy Res. 2006, 73, 492–498. [CrossRef] [PubMed] 88. Olivares, M.; Diaz-Ropero, M.A.; Gomez, N.; Lara-Villoslada, F.; Sierra, S.; Maldonado, J.A.; Martin, R.; Lopez-Huertas, E.; Rodriguez, J.M.; Xaus, J. Oral administration of two probiotic strains, Lactobacillus gasseri CECT5714 and Lactobacillus coryniformis CECT5711, enhances the intestinal function of healthy adults. Int. J. Food Microbiol. 2006, 107, 104–111. [CrossRef][PubMed] 89. Alvaro, E.; Andrieux, C.; Rochet, V.; Rigottier-Gois, L.; Leprcq, P.; Sutren, M.; Galan, P.; Duval, Y.; Juste, C.; Dore, J. Composition and metabolism of the intestinal microbiota in consumers and non-consumers of yogurt. Br. J. Nutr. 2007, 97, 126–133. [CrossRef][PubMed] Nutrients 2017, 9, 1021 24 of 30

90. Larsen, N.; Vogensen, F.K.; Gøbel, R.J.; Michaelsen, K.F.; Forssten, S.D.; Lahtinen, S.J.; Jakobsen, M. Effect of Lactobacillus salivarius Ls-33 on fecal microbiota in obese adolescents. Clin. Nutr. 2013, 32, 935–940. [CrossRef][PubMed] 91. Gøbel, R.J.; Larsen, N.; Jakobsen, M.; Mølgaard, C.; Michaelsen, K.F. Probiotics to adolescents with obesity: Effects on inflammation and metabolic syndrome. J. Pediatr. Gastroenterol. Nutr. 2012, 55, 673–678. [CrossRef] [PubMed] 92. Kadooka, Y.; Sato, M.; Imaizumi, K.; Ogawa, A.; Ikuyama, K.; Akai, Y.; Okano, M.; Kagoshima, M.; Tsuchida, T. Regulation of abdominal adiposity by probiotics (Lactobacillus gasseri SBT2055) in adults with obese tendencies in a randomized controlled trial. Eur. J. Clin. Nutr. 2010, 64, 636–643. [CrossRef][PubMed] 93. Kadooka, Y.; Sato, M.; Ogawa, A.; Miyoshi, M.; Uenishi, H.; Ogawa, H.; Ikuyama, K.; Kagoshima, M.; Tsuchida, T. Effect of Lactobacillus gasseri SBT2055 in fermented milk on abdominal adiposity in adults in a randomised controlled trial. Br. J. Nutr. 2013, 110, 1696–1703. [CrossRef][PubMed] 94. Sharafedtinov, K.K.; Plotnikova, O.A.; Alexeeva, R.I.; Sentsova, T.B.; Songisepp, E.; Stsepetova, J.; Smidt, I.; Mikelsaar, M. Hypocaloric diet supplemented with probiotic cheese improves body mass index and blood pressure indices of obese hypertensive patients—A randomized double-blind placebo-controlled pilot study. Nutr. J. 2013, 12, 138. [CrossRef][PubMed] 95. Zarrati, M.; Salehi, E.; Mofid, V.; Hossein Zadeh-Attar, M.J.; Nourijelyani, K.; Bidad, K.; Shidfar, F. Relationship between probiotic consumption and IL-10 and IL-17 secreted by PBMCs in overweight and obese people. Iran. J. Allergy Asthma Immunol. 2013, 12, 404–406. [PubMed] 96. Zarrati, M.; Salehi, E.; Nourijelyani, K.; Mofid, V.; Zadeh, M.J.; Najafi, F.; Ghaflati, Z.; Bidad, K.; Chamari, M.; Karimi, M.; et al. Effects of probiotic yogurt on fat distribution and gene expression of proinflammatory factors in peripheral blood mononuclear cells in overweight and obese people with or without weight-loss diet. J. Am. Coll. Nutr. 2014, 33, 417–425. [CrossRef][PubMed] 97. Zarrati, M.; Shidfar, F.; Nourijelyani, K.; Mofid, V.; Hossein zadeh-Attar, M.J.; Bidad, K.; Najafi, F.; Gheflati, Z.; Chamari, M.; Salehi, E. Lactobacillus acidophilus La5, Bifidobacterium BB12, and Lactobacillus casei DN001 modulate gene expression of subset specific transcription factors and cytokines in peripheral blood mononuclear cells of obese and overweight people. Biofactors 2013, 39, 633–643. [CrossRef][PubMed] 98. Agerholm-Larsen, L.; Raben, A.; Haulrik, N.; Hansen, A.S.; Manders, M.; Astrup, A. Effect of 8 week intake of probiotic milk products on risk factors for cardiovascular diseases. Eur. J. Clin. Nutr. 2000, 54, 288–297. [CrossRef][PubMed] 99. Rajkumar, H.; Mahmood, N.; Kumar, M.; Varikuti, S.R.; Challa, H.R.; Myakala, S.P. Effect of probiotic (VSL#3) and W-3 on lipid profile, insulin sensitivity, inflammatory markers, and gut colonization in overweight adults: A randomized, controlled trial. Mediat. Inflamm. 2014, 2014, 1–8. 100. Brahe, L.K.; Le Chatelier, E.; Prifti, E.; Pons, N.; Kennedy, S.; Blædel, T.; Håkansson, J.; Dalsgaard, T.K.; Hansen, T.; Pedersen, O. Dietary modulation of the gut microbiota—A randomised controlled trial in obese postmenopausal women. Br. J. Nutr. 2015, 114, 406–417. [CrossRef][PubMed] 101. Ivey, K.L.; Hodgson, J.M.; Kerr, D.A.; Thompson, P.L.; Stojceski, B.; Prince, R.L. The effect of yoghurt and its probiotics on blood pressure and serum lipid profile; a randomised controlled trial. Nutr. Metab. Cardiovasc. Dis. 2015, 25, 46–51. [CrossRef][PubMed] 102. Leber, B.; Tripolt, N.J.; Blattl, D.; Eder, M.; Wascher, T.C.; Pieber, T.R.; Stauber, R.; Sourij, H.; Oettl, K.; Stadlbauer, V. The influence of probiotic supplementation on gut permeability in patients with metabolic syndrome: An open label, randomized pilot study. Eur. J. Clin. Nutr. 2012, 66, 1110–1115. [CrossRef] [PubMed] 103. Tripolt, N.J.; Leber, B.; Blattl, D.; Eder, M.; Wonisch, W.; Scharnagl, H.; Stojakovic, T.; Obermayer-Pietsch, B.; Wascher, T.C.; Pieber, T.R.; et al. Effect of supplementation with Lactobacillus casei Shirota on insulin sensitivity, β-cell function, and markers of endothelial function and inflammation in subjects with metabolic syndrome—A pilot study. J. Dairy Sci. 2013, 96, 89–95. [CrossRef][PubMed] 104. Barreto, F.M.; Colado Simão, A.N.; Morimoto, H.K.; Batisti Lozovoy, M.A.; Dichi, I.; da Silva Miglioranza, H.L. Beneficial effects of Lactobacillus plantarum on glycemia and homocysteine levels in postmenopausal women with metabolic syndrome. Nutrition 2014, 30, 939–942. [CrossRef][PubMed] 105. Hariri, M.; Salehi, R.; Feizi, A.; Mirlohi, M.; Ghiasvand, R.; Habibi, N. A randomized, double-blind, placebo-controlled, clinical trial on probiotic soy milk and soy milk: Effects on epigenetics and oxidative stress in patients with type II diabetes. Genes Nutr. 2015, 10, 52. [CrossRef][PubMed] Nutrients 2017, 9, 1021 25 of 30

106. Tonucci, L.B.; Olbrich Dos Santos, K.M.; Licursi de Oliveira, L.; Rocha Ribeiro, S.M.; Duarte Martino, H.S. Clinical application of probiotics in type 2 diabetes mellitus: A randomized, double-blind, placebo-controlled study. Clin. Nutr. 2015, 36, 85–92. [CrossRef][PubMed] 107. Mohamadshahi, M.; Veissi, M.; Haidari, F.; Javid, A.Z.; Mohammadi, F.; Shirbeigi, E. Effects of probiotic yogurt consumption on lipid profile in type 2 diabetic patients: A randomized controlled clinical trial. J. Res. Med. Sci. 2014, 19, 531–536. [PubMed] 108. Ejtahed, H.S.; Mohtadi-Nia, J.; Homayouni-Rad, A.; Niafar, M.; Asghari-Jafarabadi, M.; Mofid, V. Probiotic yogurt improves antioxidant status in type 2 diabetic patients. Nutrition 2012, 28, 539–543. [CrossRef] [PubMed] 109. Ejtahed, H.S.; Mohtadi-Nia, J.; Homayouni-Rad, A.; Niafar, M.; Asghari-Jafarabadi, M.; Mofid, V.; Akbarian-Moghari, A. Effect of probiotic yogurt containing Lactobacillus acidophilus and Bifidobacterium lactis on lipid profile in individuals with type 2 diabetes mellitus. J. Dairy Sci. 2011, 94, 3288–3294. [CrossRef] [PubMed] 110. Andreasen, A.S.; Larsen, N.; Pedersen-Skovsgaard, T.; Berg, R.M.; Møller, K.; Svendsen, K.D.; Jakobsen, M.; Pedersen, B.K. Effects of Lactobacillus acidophilus NCFM on insulin sensitivity and the systemic inflammatory response in human subjects. Br. J. Nutr. 2010, 104, 1831–1838. [CrossRef][PubMed] 111. Vajro, P.; Mandato, C.; Licenziati, M.R.; Franzese, A.; Vitale, D.F.; Lenta, S.; Caropreso, M.; Vallone, G.; Meli, R. Effects of Lactobacillus rhamnosus strain GG in pediatric obesity-related liver disease. J. Gastroenterol. Nutr. 2011, 52, 740–743. [CrossRef][PubMed] 112. Aller, R.; de Luis, D.A.; Izaola, O.; Conde, R.; Gonzalez Sagrado, M.; Primo, D.; de la Fuente, B.; Gonzalez, J. Effect of a probiotic on liver aminotransferases in nonalcoholic fatty liver disease patients: A double blind randomized clinical trial. Eur. Rev. Med. Pharmacol. Sci. 2011, 15, 1090–1095. [PubMed] 113. Nabavi, S.; Rafraf, M.; Somi, M.H.; Homayouni-Rad, A.; Asghari-Jafarabadi, M. Effects of probiotic yogurt consumption on metabolic factors in individuals with nonalcoholic fatty liver disease. J. Dairy Sci. 2014, 97, 7386–7393. [CrossRef][PubMed] 114. Alisi, A.; Bedogni, G.; Baviera, G.; Giorgio, V.; Porro, E.; Paris, C.; Giammaria, P.; Reali, L.; Anania, F.; Nobili, V. Randomised clinical trial: The beneficial effects of VLS_3 in obese children with non-alcoholic steatohepatitis. Aliment. Pharmacol. Ther. 2014, 39, 1276–1285. [CrossRef][PubMed] 115. Wang, K.Y.; Li, S.N.; Liu, C.S.; Perng, D.S.; Su, Y.C.; Wu, D.C.; Jan, C.M.; Lai, C.H.; Wang, T.N.; Wang, W.M. Effects of ingesting Lactobacillus- and Bifidobacterium-containing yogurt in subjects with colonized Helicobacter pylori. Am. J. Clin. Nutr. 2005, 81, 939–940. 116. Cats, A.; Kuipers, E.J.; Bosschaert, M.A.; Pot, R.G.; Vandenbroucke-Grauls, C.M.; Kusters, J.G. Effect of frequent consumption of Lactobacillus casei—Containing milk drink in Helicobacter pylori-colonized subjects. Aliment. Pharmacol. Ther. 2003, 17, 429–435. [CrossRef][PubMed] 117. Kotowska, M.; Albrecht, P.; Szajewska, H. Saccharomyces boulardii in the prevention of antibiotic-associated diarrhoea in children: A randomized double-blind placebo-controlled trial. Aliment. Pharmacol. Ther. 2005, 21, 583–590. [CrossRef][PubMed] 118. Sood, A.; Midha, V.; Makharia, G.K.; Ahuja, V.; Singal, D.; Goswami, P.; Tandon, R.K. The probiotic preparation, VSL#3 induces remission in patients with mild-to-moderately active ulcerative colitis. Clin. Gastroenterol. Hepatol. 2009, 7, 1202–1209. [CrossRef][PubMed] 119. Niv, E.; Naftali, T.; Hallak, R.; Vaisman, N. The efficacy of Lactobacillus reuteri ATCC 55730 in the treatment of patients with irritable bowel syndrome–A double blind, placebo-controlled, randomized study. Clin. Nutr. 2005, 24, 925–931. [CrossRef][PubMed] 120. Wickens, K.; Black, P.N.; Stanley, T.V.; Mitchell, E.; Fitzharris, P.; Tannock, G.W.; Purdie, G.; Crane, J. A differential effect of 2 probiotics in the prevention of eczema and atopy: A double-blind, randomized, placebo-controlled trial. J. Allergy Clin. Immunol. 2008, 122, 788–794. [CrossRef][PubMed] 121. Weston, S.; Halbert, A.; Richmond, P.; Prescott, S.L. Effects of probiotics on atopic dermatitis: A randomised controlled trial. Arch. Dis. Child. 2005, 90, 892–897. [CrossRef][PubMed] 122. Niers, L.; Martin, R.; Rijkers, G.; Sengers, F.; Timmerman, H.; van Uden, N.; Smidt, H.; Kimpen, J.; Hoekstra, M. The effects of selected probiotic strains on the development of eczema (the PandA study). Allergy 2009, 64, 1349–1358. [CrossRef][PubMed] 123. Gøbel, R.J.; Larsen, N.; Mølgaard, C.; Jakobsen, M.; Michaelsen, K.F. Probiotics to young children with atopic dermatitis: A randomized placebo-controlled trial. Int. J. Probiotics Prebiotics 2010, 5, 53–60. Nutrients 2017, 9, 1021 26 of 30

124. Hertzler, S.R.; Clancy, S.M. Kefir improves lactose digestion and tolerance in adults with lactose maldigestion. J. Am. Dietetic Assoc. 2003, 103, 582–587. [CrossRef][PubMed] 125. Roškar, I.; Švigelj, K.; Štempelj, M.; Volfand, J.; Štabuc, B.; Malovrh, Š.; Rogelj, I. Effects of a probiotic product containing Bifidobacterium animalis subsp. animalis IM386 and Lactobacillus plantarum MP2026 in lactose intolerant individuals: Randomized, placebo-controlled clinical trial. J. Funct. Foods 2017, 35, 1–8. [CrossRef] 126. Liu, Z.; Qin, H.; Yang, Z.; Xia, Y.; Liu, W.; Yang, J.; Jiang, Y.; Zhang, H.; Wang, Y.; Zheng, Q. Randomized clinical trial: The effects of perioperative probiotic treatment on barrier function and post-operative infectious complications in colorectal cancer surgery—A double-blind study. Aliment. Pharmacol. Ther. 2011, 33, 50–63. [CrossRef][PubMed] 127. Chitapanarux, I.; Chitapanarux, T.; Traisathit, P.; Kudumpee, S.; Tharavichitkul, E.; Lorvidhaya, V. Randomized controlled trial of live Lactobacillus acidophilus plus Bifidobacterium bifidum in prophylaxis of diarrhea during radiotherapy in cervical cancer patients. Radiat. Oncol. 2010, 5, 31. [CrossRef][PubMed] 128. Österlund, P.; Ruotsalainen, T.; Korpela, R.; Saxelin, M.; Ollus, A.; Valta, P.; Kouri, M.; Elomaa, I.; Joensuu, H. Lactobacillus supplementation for diarrhoea related to chemotherapy of colorectal cancer: A randomised study. Br. J. Cancer 2007, 97, 1028–1034. [CrossRef][PubMed] 129. Chung, W.S.F.; Walker, A.W.; Louis, P.; Parkhill, J.; Vermeiren, J.; Bosscher, D.; Duncan, S.H.; Flint, H.J. Modulation of the human gut microbiota by dietary fibres occurs at the species level. BMC Biol. 2016, 14, 1–13. [CrossRef][PubMed] 130. Crittenden, R.; Playne, M.J. Nutrition News. Facts and functions of prebiotics, probiotics and synbiotics. In Handbook of Probiotics and Prebiotics; Lee, Y.K., Salminen, S., Eds.; Wiley-Interscience, Kansas State University: Hoboken, NJ, USA; Manhattan, KS, USA, 2008; pp. 535–582. 131. Jakubczyk, E.; Kosikowska, M. Nowa generacja mlecznych produktów fermentowanych z udziałem probiotyków i prebiotyków, produkty synbiotyczne. Prz. Mlecz. 2000, 12, 397–400. 132. Wang, Y. Prebiotics: Present and future in food science and technology. Food Res. Int. 2009, 42, 8–12. [CrossRef] 133. Maccfarlane, G.T.; Steed, H.; Maccfarlane, S. Bacterial metabolism and health-related effects of galacto-oligosaccharides and other prebiotics. J. Appl. Microbiol. 2008, 104, 305–344. [CrossRef][PubMed] 134. Crittenden, R.; Playne, M.J. Prebiotics. In Handbook of Probiotics and Prebiotics; Lee, Y.K., Salminen, S., Eds.; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2009; pp. 535–561. 135. Huebner, J.; Wehling, R.L.; Parkhurst, A.; Hutkins, R.W. Effect of processing conditions on the prebiotic activity of commercial prebiotics. Int. Dairy J. 2008, 18, 287–293. [CrossRef] 136. Ze, X.; Duncan, S.H.; Louis, P.; Flint, H.J. Ruminococcus bromii is a keystone species for the degradation of in the human colon. ISME J. 2012, 6, 1535–1543. [CrossRef][PubMed] 137. Angelakis, E. Weight gain by gut microbiota manipulation in productive animals. Microb. Pathog. 2017, 106, 162–170. [CrossRef][PubMed] 138. Sli´ zewska,˙ K.; Nowak, A.; Barczy´nska,R.; Libudzisz, Z. Prebiotyki—Definicja, wła´sciwo´scii zastosowanie w przemy´sle. Zywno´s´cNauka˙ Technologia Jako´s´c 2013, 1, 5–20. 139. Sivieri, K.; Morales, M.L.V.; Saad, S.M.I.; Adorno, M.A.T.; Sakamoto, I.K.; Rossi, E.A. Prebiotic effect of in the simulator of the human intestinal microbial ecosystem (SHIME (R) Model). J. Med. Food 2014, 17, 894–901. [CrossRef][PubMed] 140. Van Den Abbeele, P.; Venema, K.; van de Wiele, T.; Verstraete, W.; Possemiers, S. Different human gut models reveal the distinct fermentation patterns of arabinoxylan versus inulin. J. Agric. Food Chem. 2013, 61, 9819–9827. [CrossRef][PubMed] 141. Ouwehand, A.; Derrien, M.; de Vos, W.; Tiihonen, K.; Rautonen, N. Prebiotics and other microbial substrates for gut functionality. Curr. Biol. 2005, 16, 212–217. [CrossRef][PubMed] 142. Patterson, J.A.; Burkholder, K.M. Application of prebiotics and probiotics in poultry production. Poult. Sci. 2003, 82, 627–631. [CrossRef][PubMed] 143. Annison, G.; Illman, R.; Topping, D. Acetylated, propionylated or butyrylated raise large bowel short-chain fatty acids preferentially when fed to rats. J. Nutr. 2003, 133, 3523–3528. [PubMed] 144. Baurhoo, B.; Letellier, A.; Zhao, X.; Ruiz-Feria, C.A. Cecal populations of Lactobacilli and Bifidobacteria and Escherichia coli after In Vivo Escherichia coli challenge in birds fed diets with purified lignin or mannanoligo-saccharides. Poult. Sci. 2007, 86, 2509–2516. [CrossRef][PubMed] Nutrients 2017, 9, 1021 27 of 30

145. Olveira, G.; González-Molero, I. An update on probiotics, prebiotics and symbiotics in clinical nutrition. Endocrinol. Nutr. 2016, 63, 482–494. [CrossRef][PubMed] 146. Sáez-Lara, M.J.; Robles-Sanchez, C.; Ruiz-Ojeda, F.J.; Plaza-Diaz, J.; Gil, A. Effects of probiotics and synbiotics on obesity, insulin resistance syndrome, type 2 diabetes and non-alcoholic fatty liver disease: A review of human clinical trails. Int. J. Mol. Sci. 2016, 17, 928. [CrossRef][PubMed] 147. Van Loo, J.; Clune, Y.; Bennett, M.; Collins, J.K. The SYNCAN project: Goals, set-up, first results and settings of the human intervention study. Br. J. Nutr. 2005, 93, S91–S98. [PubMed] 148. Schiffrin, E.J.; Kumar, V.B.; Brown, C.; Hager, C.; Van’t Hof, M.A.; Morley, J.E.; Guigoz, Y. Systemic inflammatory markers in older persons: The effect of oral nutritional supplementation with prebiotics. J. Nutr. Health Aging 2007, 11, 475–479. [PubMed] 149. Vulevic, J.; Drakoularakou, A.; Yaqoob, P.; Tzortzis, G.; Gibson, G.R. Modulation of the fecal microflora profile and immune function by a novel transgalactooligosaccharide mixture (B-GOS) in healthy elderly volunteers. Am. J. Clin. Nutr. 2008, 88, 1438–1446. [PubMed] 150. Schley, P.D.; Field, C.J. The immune-enhancing effects of dietary fibres and prebiotics. Br. J. Nutr. 2002, 87, S221–S230. [CrossRef][PubMed] 151. Grajek, W.; Olejnik, A.; Sip, A. Probiotics, prebiotics and antioxidants as functional foods. Acta Biochim. Pol. 2005, 52, 665–671. [PubMed] 152. Gibson, G.R.; Wang, X. Regulatory effects of bifidobacteria on the growth of other colonic bacteria. J. Appl. Microbiol. 1994, 77, 412–420. [CrossRef] 153. Bovee-Oudenhoven, I.M.J.; Termont, D.S.; Heidt, P.J.; van der Meer, R. Increasing the intestinal resistance of rats to the invasive pathogen Salmonella enteritidis: Additive effects of dietary lactulose and calcium. Gut 1997, 40, 497–504. [CrossRef][PubMed] 154. De Preter, V.; Hamer, H.M.; Windey, K.; Verbeke, K. The impact of pre- and/or probiotics on human colonic metabolism: Does it affect human health? Mol. Nutr. Food Res. 2011, 55, 46–57. [CrossRef][PubMed] 155. Demigné, C.; Jacobs, H.; Moundras, C.; Davicco, M.J.; Horcajada, M.N.; Bernalier, A.; Coxam, V. Comparison of native or reformulated chicory fructans, or non-purified chicory, on rat cecal fermentation and mineral metabolism. Eur. J. Nutr. 2008, 47, 366–374. [CrossRef][PubMed] 156. Mojka, K. Probiotyki, prebiotyki i synbiotyki—Charakterystyka i funkcje. Probl. Hig. Epidemiol. 2014, 95, 541–549. 157. Socha, P.; Stolarczyk, M.; Socha, J. Wpływ probiotyków i prebiotyków na gospodark˛e lipidow ˛a. Pediatr. Współcz. Gastroenterol. Hepatol. Zyw.˙ Dziecka 2002, 4, 85–88. 158. Asahara, T.; Nomoto, K.; Shimizu, K.; Watanuki, M.; Tanaka, R. Increased resistance of mice to Salmonella enteritica serovar Typhymurium infection by synbiotic administration of bifidobacteria and transgalactosylated-oligosaccharides. J. Appl. Microbiol. 2001, 91, 985–996. [CrossRef][PubMed] 159. Buddington, K.K.; Danohoo, J.B.; Buddington, R.K. Dietary oligofructose and inulin protect mice from enteric and systemic pathogens and tumour inducers. J. Nutr. 2002, 132, 472–477. [PubMed] 160. Cummings, J.H.; Macfarlane, G.T. Gastrointestinal effects of prebiotics. Br. J. Nutr. 2002, 87, 145–151. [CrossRef] 161. Scheppach, W.; Weiler, F. The butyrate story: Old wine in new bottles? Butyrate appears to be essential for a wide range of intestinal mucosal health benefits; however, the mechanisms behind this remain to be determined. Curr. Opin. Clin. Nutr. Metab. Care 2004, 7, 563–567. [CrossRef][PubMed] 162. Kim, Y.S.; Tsa, O.D.; Morita, A.; Bella, A. Effect of sodium butyrate and three human colorectal adenocarcinoma cell lines in culture. Falk Symp. 1982, 31, 317–323. 163. Munjal, U.; Glei, M.; Pool-Zobel, B.L.; Scharlau, D. Fermentation products of inulin-type fructans reduce proliferation and induce apoptosis in human colon tumour cells of different stages of carcinogenesis. Br. J. Nutr. 2009, 27, 1–9. [CrossRef][PubMed] 164. Verghese, M.; Rao, D.R.; Chawan, C.B.; Williams, L.L.; Shackelford, L. Dietary inulin suppresses azoxymethane-induced aberrant crypt foci and colon tumors at the promotion stage in young Fisher 344 rats. J. Nutr. 2002, 132, 2809–2813. [PubMed] 165. Taper, H.S.; Roberfroid, M.B. Inulin/Oligofructose and anticancer therapy. Br. J. Nutr. 2002, 87, 283–286. [CrossRef] Nutrients 2017, 9, 1021 28 of 30

166. Parnell, J.A.; Reimer, R.A. Weight loss during oligofructose supplementation is associated with decreased ghrelin and increased peptide YY in overweight and obese adults. Am. J. Clin. Nutr. 2009, 89, 1751–1759. [CrossRef][PubMed] 167. Luo, J.; van Yperselle, M.; Rizkalla, S.W.; Rossi, F.; Bornet, F.R.; Slama, G. Chronic consumption of short-chain fructooligosaccharides does not affect basal hepatic glucose production or insulin resistance in type 2 diabetics. J. Nutr. 2000, 130, 1572–1577. [PubMed] 168. Lu, Z.X.; Walker, K.Z.; Muir, J.G.; O’Dea, K. Arabinoxylan fibre improves metabolic control in people with type II diabetes. Eur. J. Clin. Nutr. 2004, 58, 621–628. [CrossRef][PubMed] 169. Garcia, A.L.; Otto, B.; Reich, S.C. Arabinoxylan consumption decreases postprandial serum glucose, serum insulin and plasma total ghrelin response in subjects with impaired glucose tolerance. Eur. J. Clin. Nutr. 2007, 61, 334–341. [CrossRef][PubMed] 170. Daubioul, C.A.; Horsmans, Y.; Lambert, P.; Danse, E.; Delzenne, N.M. Effects of oligofructose on glucose and lipid metabolism in patients with nonalcoholic steatohepatitis: Results of a pilot study. Eur. J. Clin. Nutr. 2005, 59, 723–726. [CrossRef][PubMed] 171. Bruzzese, E.; Volpicelli, M.; Salvini, F.; Bisceglia, M.; Lionetti, P.; Cinquetti, M.; Iacono, G.; Guarino, A. Early administration of GOS/FOS prevents intestinal and respiratory infections in infants. J. Pediatr. Gastroenterol. Nutr. 2006, 42, E95. 172. Costalos, C.; Kapiki, A.; Apostolou, M.; Papathoma, E. The effect of a prebiotic supplemented formula on growth and stool microbiology of term infants. Early Hum. Dev. 2008, 84, 45–49. [CrossRef][PubMed] 173. Scholtens, P.A.; Alliet, P.; Raes, M.; Alles, M.S.; Kroes, H.; Boehm, G.; Knippels, L.M.; Knol, J.; Vandenplas, Y. Fecal secretory immunoglobulin A is increased in healthy infants who receive a formula with short-chain galacto-oligosaccharides and long-chain fructo-oligosaccharides. J. Nutr. 2008, 138, 1141–1147. [PubMed] 174. Welters, C.F.; Heineman, E.; Thunnissen, F.B.; van den Bogaard, A.E.; Soeters, P.B.; Baeten, C.G. Effect of dietary inulin supplementation on inflammation of pouch mucosa in patients with an ileal pouch-anal anastomosis. Dis. Colon Rectum 2002, 45, 621–627. [CrossRef][PubMed] 175. Lindsay, J.O.; Whelan, K.; Stagg, A.J.; Gobin, P.; Al-Hassi, H.O.; Rayment, N.; Forbes, A. Clinical, microbiological, and immunological effects of fructo- in patients with Crohn’s disease. Gut 2006, 55, 348–355. [CrossRef][PubMed] 176. Moro, G.; Arslanoglu, S.; Stahl, B.; Jelinek, J.; Wahn, U.; Boehm, G. A mixture of prebiotic oligosaccharides reduces the incidence of atopic dermatitis during the first six months of age. Arch. Dis. Child. 2006, 91, 814–819. [CrossRef][PubMed] 177. Arslanoglu, S.; Moro, G.E.; Boehm, G. Early supplementation of prebiotic oligosaccharides protects formula-fed infants against infections during the first 6 months of life. J. Nutr. 2007, 137, 2420–2424. [PubMed] 178. Azcarate-Peril, M.A.; Ritter, A.J.; Savaiano, D.; Monteagudo-Mera, A.; Anderson, C.; Magness, S.T.; Klaenhammer, T.R. Impact of short-chain galactooligosaccharides on the gut of lactose-intolerant individuals. Proc. Natl. Acad. Sci. USA 2017, 114, E367–E375. [CrossRef][PubMed] 179. Gourbeyre, P.; Denery, S.; Bodinier, M. Probiotics, prebiotics, and synbiotics: Impact on the gut and allergic reactions. J. Leukoc. Biol. 2011, 89, 685–695. [CrossRef][PubMed] 180. De Vrese, M.; Schrezenmeir, J. Probiotics, prebiotics and synbiotics. In Food Biotechnology, Advances in Biochemical Engineering/Biotechnology; Stahl, U., Donalies, U.E.B., Nevoigt, E., Eds.; Springer: Berlin, Germany, 2008; pp. 1–66. [CrossRef] 181. Scavuzzi, B.M.; Henrique, F.C.; Miglioranza, L.H.S.; Simão, A.N.C.; Dichi, I. Impact of prebiotics, probiotics and synbiotics on components of the metabolic syndrome. Ann. Nutr. Disord. Ther. 2014, 1, 1009. 182. Hamasalim, H.J. Synbiotic as feed additives relating to animal health and performance. Adv. Microbiol. 2016, 6, 288–302. [CrossRef] 183. Sekhon, B.S.; Jairath, S. Prebiotics, probiotics and synbiotics: An overview. J. Pharm. Educ. Res. 2010, 1, 13–36. 184. Manigandan, T.; Mangaiyarkarasi, S.P.; Hemaltha, R.; Hemaltha, V.T.; Murali, N.P. Probiotics, prebiotics and synbiotics—A review. Biomed. Pharmacol. J. 2012, 5, 295–304. [CrossRef] 185. Pandey, K.R.; Naik, S.R.; Babu, V.V. Probiotics, prebiotics and synbiotics—A review. J. Food. Sci. Technol. 2015, 52, 7577–7587. [CrossRef][PubMed] 186. Zhang, M.M.; Cheng, J.Q.; Lu, Y.R.; Yi, Z.H.; Yang, P.; Wu, X.T. Use of pre-, pro-and synbiotics in patients with acute pancreatitis: A meta-analysis. World J. Gastroenterol. 2010, 16, 3970. [CrossRef][PubMed] Nutrients 2017, 9, 1021 29 of 30

187. Wong, V.W.; Won, G.L.; Chim, A.M.; Chu, W.C.; Yeung, D.K.; Li, K.C.; Chan, L. Treatment of nonalcoholic steatohepatitis with probiotics. A proof-of-concept study. Ann. Hepatol. 2013, 12, 256–262. [PubMed] 188. Eslamparast, T.; Poustchi, H.; Zamani, F.; Sharafkhah, M.; Malekzadeh, R.; Hetmatdoost, A. Synbiotic supplementation in nonalcoholic fatty liver disease: A randomized, double-blind, placebo-controlled pilot study. Am. J. Clin. Nutr. 2014, 99, 535–542. [CrossRef][PubMed] 189. Pathmakanthan, S.; Walsh, M.; Bengmark, S. Efficacy and Tolerability Treating Acute Distal Ulcerative Colitis with Synbiotic Enemas: A Pilot Trial (Abstract); United European Gastroenterology Week: Geneva, Switzerland, 2002. 190.P érez-Conesa, D.; López, G.; Abellán, P.; Ros, G. Bioavailability of calcium, magnesium and phosphorus in rats fed probiotic, prebiotic and synbiotic powder follow-up infant formulas and their effect on physiological and nutritional parameters. J. Sci. Food Agric. 2006, 86, 2327–2336. [CrossRef] 191. Danq, D.; Zhou, W.; Lun, Z.J.; Mu, X.; Wanq, D.X.; Wu, H. Meta-analysis of probiotics and/or prebiotics for the prevention of eczema. J. Int. Med. Res. 2013, 41, 1426–1436. [CrossRef][PubMed] 192. Rafter, J.; Bennett, M.; Caderni, G.; Clune, I.; Hughes, R.; Karlsson, P.C.; Klinder, A.; O’Riordan, M.; O’Sullivan, G.; Pool-Zobel, B.; et al. Dietary synbiotics reduce cancer risk factors in polypectomized and colon cancer patients. Am. J. Clin. Nutr. 2007, 85, 488–496. [PubMed] 193. Sanchez, M.; Darimont, C.; Drapeau, V.; Emady-Azar, S.; Lepage, M.; Rezzonico, E.; Ngom-Bru, C.; Berger, B.; Philippe, L.; Ammon-Zuffrey, C.; et al. Effect of Lactobacillus rhamnosus CGMCC1.3724 supplementation on weight loss and maintenance in obese men and women. Br. J. Nutr. 2014, 111, 1507–1519. [CrossRef] [PubMed] 194. Safavi, M.; Farajian, S.; Kelishadi, R.; Mirlohi, M.; Hashemipour, M. The effects of synbiotic supplementation on some cardio metabolic risk factors in overweight and obese children: A randomized triple-masked controlled trial. Int. J. Food Sci. Nutr. 2013, 64, 687–693. [CrossRef][PubMed] 195. Ipar, N.; Aydogdu, S.D.; Yildirim, G.K.; Inal, M.; Gies, I.; Vandenplas, Y.; Dinleyici, E.C. Effects of symbiotic on anthropometry, lipid profile and oxidative stress in obese children. Benef. Microbes 2015, 6, 775–781. [CrossRef][PubMed] 196. Eslamparast, T.; Zamani, F.; Hekmatdoost, A.; Sharafkhah, M.; Eghtesad, S.; Malekzadeh, R.; Poustchi, H. Effects of synbiotic supplementation on insulin resistance in subjects with the metabolic syndrome: A randomised, double-blind, placebo-controlled pilot study. Br. J. Nutr. 2014, 112, 438–445. [CrossRef] [PubMed] 197. Asemi, Z.; Zohreh, Z.; Shakeri, H.; Sima-sadat Sabihi, S.S.; Esmaillzadeh, A. Effect of multispecies probiotic supplements on metabolic profiles, hs-CRP, and oxidative stress in patients with type 2 diabetes. Ann. Nutr. Metab. 2013, 63, 1–9. [CrossRef][PubMed] 198. Tajadadi-Ebrahimi, M.; Bahmani, F.; Shakeri, H.; Hadaegh, H.; Hijijafari, M.; Abedi, F.; Asemi, Z. Effects of daily consumption of synbiotic bread on insulin metabolism and serum high-sensitivity C-reactive protein among diabetic patients: A double-blind, randomized, controlled clinical trial. Ann. Nutr. Metab. 2014, 65, 34–41. [CrossRef][PubMed] 199. Shakeri, H.; Hadaegh, H.; Abedi, F.; Tajabadi-Ebrahimi, M.; Mazroii, N.; Ghandi, Y.; Asemi, Z. Consumption of synbiotic bread decreases triacylglycerol and VLDL levels while increasing HDL levels in serum from patients with type-2 diabetes. Lipids 2014, 49, 695–701. [CrossRef][PubMed] 200. Moroti, C.; Souza Magri, L.F.; Rezende-Costa, M.; Cavallini, D.; Sivieri, K. Effect of the consumption of a new symbiotic shake on glycemia and cholesterol levels in elderly people with type 2 diabetes mellitus. Lipids Health Dis. 2012, 11, 29. [CrossRef][PubMed] 201. Šmid, A.; Strniša, L.; Bajc, K.; Vuji´c-Podlipec,D.; BogoviˇcMatijaši´c,B.; Rogelj, I. Randomized clinical trial: The effect of fermented milk with the probiotic cultures Lactobacillus acidophilus La-5® and Bifidobacterium BB-12® and Beneo dietary fibres on health-related quality of life and the symptoms of irritable bowel syndrome in adults. J. Funct. Foods 2016, 24, 549–557. [CrossRef] 202. Ustundag, G.H.; Altuntas, H.; Soysal, Y.D.; Kokturk, F. The effects of synbiotic “Bifidobacterium lactis B94 plus Inulin” addition on standard triple therapy of Helicobacter pylori eradication in children. Can. J. Gastroenterol. Hepatol. 2017, 2017, 8130596. [CrossRef][PubMed] 203. Fujimori, S.; Gudis, K.; Mitsui, K.; Seo, T.; Yonezawa, M.; Tanaka, S.; Tatsuguchi, A.; Sakamoto, C. A randomized controlled trial on the efficacy of synbiotic versus probiotic or prebiotic treatment to improve the quality of life in patients with ulcerative colitis. Nutrition 2009, 25, 520–525. [CrossRef][PubMed] Nutrients 2017, 9, 1021 30 of 30

204. Van der Aa, L.B.; Heymans, H.S.A.; van Aalderen, W.M.C.; Sillevis Smitt, J.H.; Knol, J.; Ben Amor, K.; Goossens, D.A.; Sprikkelman, A.B. Effect of a new synbiotic mixture on atopic dermatitis in infants: A randomized controlled trial. Clin. Exp. Allergy 2010, 40, 795–804. [CrossRef][PubMed] 205. Farid, R.; Ahanchian, H.; Jabbari, F.; Moghiman, T. Effect of a new synbiotic mixture on atopic dermatitis in children: A randomized-controlled trial. Iran. J. Pediatr. 2011, 21, 225–230. [PubMed] 206. Post, J. Acceptability and Feasibility of Probiotic and Prebiotic Supplementation in Alleviating Symptoms of Lactose Maldigestion in Lactose Intolerant Subjects; Florida State University: Tallahassee, FL, USA, 2013. 207. Nowak, A.; Sli´ zewska,˙ K.; Libudzisz, Z. Probiotyki—Historia i mechanizmy działania. Zywno´s´cNauka˙ Technologia Jako´s´c 2010, 4, 5–19. 208. Bomba, A.; Nemcova, R.; Mudronova, D.; Guba, P. The possibilities of potentiating the efficacy of probiotics. Trends Food Sci. Technol. 2002, 13, 121–126. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).